structure-function insights into the biochemical

178
Structure-Function Insights into the Biochemical Properties and Bilayer Interactions of the Saposin-like Domain of Plant Aspartic Proteases by Brian C. Bryksa A Thesis Presented to The University of Guelph In partial fulfilment of requirements for the degree of Doctor of Philosophy in Food Science Guelph, Ontario, Canada © Brian C. Bryksa, December, 2016

Upload: others

Post on 08-Feb-2022

1 views

Category:

Documents


0 download

TRANSCRIPT

Structure-Function Insights into the Biochemical Properties and Bilayer

Interactions of the Saposin-like Domain of Plant Aspartic Proteases

by

Brian C. Bryksa

A Thesis

Presented to

The University of Guelph

In partial fulfilment of requirements

for the degree of

Doctor of Philosophy

in

Food Science

Guelph, Ontario, Canada

© Brian C. Bryksa, December, 2016

ABSTRACT

STRUCTURE-FUNCTION INSIGHTS INTO THE BIOCHEMICAL

PROPERTIES AND BILAYER INTERACTION OF THE SAPOSIN-LIKE

DOMAIN OF PLANT ASPARTIC PROTEASES

Brian C. Bryksa Advisor:

University of Guelph, 2016 Professor R.Y. Yada

This thesis is an investigation of structure-function relationships of the saposin-like domain

of plant aspartic proteases. Many plant aspartic proteases contain an additional sequence

of approximately 100 amino acids termed the plant-specific insert which is involved in host

defense and vacuolar targeting. Similar to all other saposin-like proteins, the plant-specific

insert functions via protein-membrane interactions, however, the structural basis for such

interactions have not been studied and the nature of plant-specific insert-mediated

membrane disruption have not been characterized. This thesis presents the first

comprehensive structure-function investigation of the less-understood arm of the saposin-

like protein family, the so-called “swaposins”. Among the findings presented here are the

quaternary, tertiary and secondary structures of the plant-specific insert from Solanum

tuberosum (potato) aspartic proteinase both in terms of pH and lipid bilayer presence, the

identification of a structure in potato saposin (Ile1-Leu20) that is a universal membrane

penetrating motif based upon structural alignment, delineation of the structural basis for

the acid pH requirement for bilayer interaction (pH-sensitive dimerization) and positively

charged point of contact for anionic bilayers (Lys83) located at the C-terminal end of helix

3, a positively charged residue within an uncommon anti-bilayer motif found in some

flocculant proteins and spider silk structural proteins, among others. Atomic force

microscopy revealed that potato plant-specific insert destabilized (softened) bilayer

whereas cryo-transmission electron microscopy showed several distinct shapes induced in

LUV’s upon addition of potato saposin. Lastly, comparative characterizations of potato

plant-specific insert along with three other plant saposin-like domains from barley

(Hordeum vulgare), flowers of Cardoon thistle (Cynara cardunculus L.) and Rockcress

(Arabidopsis thaliana) were carried out revealing that reduction of the disulfide bonds of

potato swaposin caused a drastic increase in bilayer fusion rate and increase in typical

fusion product sizes. Arabidopsis swaposin with reduced cystines showed relatively minor

alterations to its fusion profile while essentially no difference to the fusogenic activities of

barley and Cardoon swaposins were discernable upon reduction of their disulfide bonds.

Taken together, implications for swaposin mechanism of action as well as future research

directions are discussed.

iii

Acknowledgments

First and foremost, I am grateful to my advisor and mentor, Dr. Rickey Yada, whose

contributions to my research endeavours and life beyond have been uniquely

transformative and enriching. Along with Dr. Yada, I am thankful for the inspiration of Dr.

Takuji Tanaka and Dr. Jong Kun Ahn, who accepted and taught me at the very beginning.

Three busy researchers offered their time, patience and attention to a second year

undergraduate without lab experience, but a willingness to volunteer. Furthermore, I am

grateful for the interest, time and input of my advisory committee members Dr. Alejandro

Marangoni, Dr. Leonid Brown and Dr. John Dutcher, as well as collaborators Dr. Marty

Kurylowicz, Dr. Prasenjit Bhaumik, Eugenia Magracheva, Dr. Alexander Zdanov

(deceased) and Dr. Alexander Wlodawer. To the many undergraduate project and summer

students whom I have taught, and who have taught me more, and group members,

coworkers and professors with whom I have had the honour of sharing knowledge and

experience, too many to name here, I offer sincere thanks. In particular, thank you to those

who assisted with the experiments described herein: Lauren Agro, Sean Filonowicz, Dref

de Moura, Phil Formusa, Kara Griffiths, Jenny Tian, Kassandra Wagner, Sean Harrington,

Michael Bregler, Doug Grahame, Dr. Frances Sharom and Dr. Massimo Marcone. I am

also grateful for the essential support, advice and comfort of friends and family. I thank

Heather, Kathryn, Emily and Megan, who have lifted and sustained me in ways that I am

incapable of understanding let alone properly describing here. Lastly, the Natural Sciences

and Engineering Research Council of Canada, and the Canada Research Chairs program

are gratefully acknowledged for funding the research described in this thesis.

iv

Table of Contents

ABSTRACT ....................................................................................................................... ii

Acknowledgments ............................................................................................................ iii

Table of Contents ............................................................................................................. iv

List of Tables ................................................................................................................... vii

List of Figures ................................................................................................................. viii

List of Abbreviations ........................................................................................................ x

Chapter 1: Introduction ................................................................................................... 1

1.1 Background ............................................................................................................... 1

1.2 Literature Review ...................................................................................................... 2

1.2.1 Plant aspartic proteinases .................................................................................... 2 1.2.2 Solanum tuberosum aspartic proteinases (StAP) ................................................ 7

1.2.3 Saposins ............................................................................................................ 10 1.2.4 Structure-Function Studies of Saposin Disulfide Bonds .................................. 12 1.2.5 Plant AP PSI-Membrane Interactions ............................................................... 14

1.3 Scientific Investigation Plan .................................................................................... 14

1.3.1 Research Questions ........................................................................................... 14 1.3.2 Objectives ......................................................................................................... 15 1.3.3 Hypotheses ........................................................................................................ 16

1.4 References ............................................................................................................... 17

Chapter 2: Structure and Mechanism of the Saposin-Like Domain of a Plant

Aspartic Protease .......................................................................................................... 24

2.1 Abstract ................................................................................................................... 24

2.2 Introduction ............................................................................................................. 25

2.3 Experimental Procedures ......................................................................................... 28

2.3.1 Materials ........................................................................................................... 28

2.3.2 Construction ...................................................................................................... 28 2.3.3 Protein Expression ............................................................................................ 29 2.3.4 Protein Purification ........................................................................................... 29 2.3.5 SDS-PAGE ....................................................................................................... 30 2.3.7 Diffraction Data Collection, Structure Solution and Refinement ..................... 31

2.3.8 Circular Dichroism (CD) Spectropolarimetry - ................................................ 31 2.3.9 Preparation of Large Unilamellar Vesicles (LUV) ........................................... 32 2.3.10 LUV Disruption Assays .................................................................................. 32 2.3.11 Atomic Force Microscopy .............................................................................. 33

2.3.12 Particle Size Determination by Light Scattering ............................................ 33

2.4 Results ..................................................................................................................... 34

2.4.1 Structure Solution and Refinement ................................................................... 34 2.4.2 Tertiary and Quaternary Structures of StAP PSI .............................................. 36

v

2.4.3 pH Dependence of Secondary Structure ........................................................... 36

2.4.4 Secondary Structure Dependence on Disulfide bonds...................................... 38 2.4.5 Membrane Disruption Activity – Vesicle Leakage .......................................... 38 2.4.6 Membrane Disruption Activity – Atomic Force Microscopy .......................... 44

2.4.7 Membrane Disruption Activity – Light Scattering ........................................... 46

2.5 Discussion ............................................................................................................... 48

2.5.1 Structural Comparison ...................................................................................... 48 2.5.2 Saposin-like Activity ........................................................................................ 48 2.5.3 Plant AP PSI-Activity ....................................................................................... 51 2.5.4 Fusogenic Mechanism ...................................................................................... 53

2.5.5 Fusase Within a Protease .................................................................................. 55

2.6 References ............................................................................................................... 57

Chapter 3: Protein Structure Insights into the Bilayer Interactions of the Saposin-

Like Domain of Solanum tuberosum Aspartic Protease ............................................ 65

3.1 Abstract ................................................................................................................... 65

3.2 Introduction ............................................................................................................. 66

3.3 Materials and Methods ............................................................................................ 69

3.3.1 Materials ........................................................................................................... 69 3.3.2 PSI Expression and Purification ....................................................................... 69

3.3.3 Preparation of Large Unilamellar Vesicles ...................................................... 71 3.3.4 Circular Dichroism Spectropolarimetry ........................................................... 71 3.3.5 LUV Disruption Assays .................................................................................... 71

3.3.6 Bilayer Fusion Assays ...................................................................................... 72

3.3.7 Sedimentation Equilibrium Analytical Centrifugation ..................................... 72 3.3.8 Tryptophan Intrinsic Fluorescence Emission Spectrometry ............................. 73 3.3.9 Cryo-Transmission Electron Microscopy ......................................................... 73

3.3.10 PSI Structure Component Peptides ................................................................. 74 3.3.11 Statistical Analyses ......................................................................................... 75

3.4 Results ..................................................................................................................... 75

3.4.1 StAP PSI Quaternary Structure in Solution ...................................................... 75 3.4.2 Secondary Structure Dependence on pH and Disulfide Bonds ........................ 78

3.4.3 pH-Dependence of PSI-induced Phospholipid Bilayer Disruption .................. 80 3.4.4 Intrinsic Tryptophan Fluorescence for PSI in the Presence of Anionic Bilayer

................................................................................................................................... 82 3.4.5 PSI Secondary Structure in the Presence of Anionic Bilayer LUVs ................ 82

3.4.6 Characterization of PSI-induced Bilayer Effects: Cryo-Transmission Electron

Microscopy ................................................................................................................ 85 3.4.7 Characterization of Bilayer Fusion Activity: Dynamic Light Scattering ......... 89

3.4.8 PSI Component Peptide-Induced Bilayer Disruption ....................................... 91 3.4.9 H3 Intrinsic Tryptophan Fluorescence in the presence of Anionic Bilayer

Vesicles ...................................................................................................................... 93 3.4.10 Secondary Structure of PSI Component Peptides .......................................... 93 3.4.11 Peptide-induced vesicle fusion ....................................................................... 96

vi

3.5 Discussion ............................................................................................................... 98

3.6 References ............................................................................................................. 112

3.7 Appendices ............................................................................................................ 122

3.7.A Considerations for using DLS to characterize vesicle fusion ........................ 122 3.7.B Stability of large unilamellar vesicles ............................................................ 122 3.7.C Considerations for comparing leakage rates at low pH values ...................... 124

Chapter 4: Comparative Structure-Function Characterization of the Saposin-Like

Domains from Potato, Barley, Cardoon Thistle and Arabidopsis Aspartic Proteases

....................................................................................................................................... 126

4.1 Abstract ................................................................................................................. 126

4.2 Introduction ........................................................................................................... 127

4.3 Experimental Procedures ....................................................................................... 129

4.3.1 Materials ......................................................................................................... 129 4.3.2 PSI Expression and Purification ..................................................................... 130 4.3.3 Preparation of Large Unilamellar Vesicles (LUV) ......................................... 130 4.3.4 Circular Dichroism Spectropolarimetry (CD) ................................................ 130

4.3.5 LUV Disruption Assays .................................................................................. 130 4.3.6 Bilayer Fusion Assays .................................................................................... 132

4.3.7 Tryptophan Intrinsic Fluorescence Spectrometry ........................................... 132

4.4 Results and Discussion .......................................................................................... 133

4.4.1 Primary Structure Comparison ....................................................................... 133 4.4.2 Bilayer Disruption and Fusion ........................................................................ 135

4.4.3 Comparison of pH-dependence of secondary structure .................................. 142

4.4.4 Intrinsic Trp fluorescence in solution and PSI-bilayer interactions ............... 143

4.5 References ............................................................................................................. 149

4.6 Appendices ............................................................................................................ 153

4.6-Appendix A Comparison of the secondary structures of StAP, phytepsin,

cardosin A and AtAP PSIs at different pH in iso-ionic buffered saline. ................. 153 4.6-Appendix B Comparison of the secondary structures of StAP, phytepsin,

cardosin A and AtAP PSIs at different pH in iso-ionic buffered saline .................. 154

Chapter 5: Concluding Discussion .............................................................................. 155

5.1 Further Considerations for PSI-Induced Bilayer Fusion ....................................... 155

5.2 Bilayer Activity Via an Isolated Swaposin Structural Region .............................. 159

5.3 Considerations for in vitro Structure-Function Studies on Bioactive Proteins ..... 161

5.4 Final Thoughts on Future Research Directions ..................................................... 162

5.5 References ............................................................................................................. 166

vii

List of Tables

Table 2.1: Data collection and refinement statistics ......................................................... 35

Table 2.2: Turnover and goodness of fit to the Michaelis-Menten model for StAP PSI-

induced vesicle leakage .................................................................................. 43

Table 2.3: Effect of StAP PSI on LUV size at pH 4.5 ...................................................... 47

Table 3.1: Examples of antimicrobial/membrane-interacting proteins that contain the

[N/Q]-[N/Q]-[A/L/I/V]-[R/K]-[N/Q] motif. ................................................ 108

Table 3.2: Proteins in the RCSB PDB databank that contain the [N/Q]-[N/Q]-[N/Q]-

[A/L/I/V]-[R/K]-[N/Q] motif ....................................................................... 109

Table 4.1: Phospholipid composition used for preparing LUVmix ................................ 131

Table 5.1: Biologically active concentrations for StAP PSI ........................................... 163

Table 5.2: Concentration ranges of antimicrobial peptides used in vitro ....................... 163

viii

List of Figures

Figure 1.1: The aspartic protease bilobal structure. ............................................................ 3

Figure 1.2: Plant aspartic protease insert. ........................................................................... 5

Figure 1.3: The crystal structure of prophytepsin (1QDM). ............................................... 5

Figure 1.4: The canonical saposin fold. .............................................................................. 6

Figure 1.5: Comparison of the two known saposin folds. ................................................ 13

Figure 2.1: The structure of StAP PSI. ............................................................................. 37

Figure 2.2: Effect of acidification on StAP PSI secondary structure ............................... 39

Figure 2.3: Importance of disulfide bonds on StAP PSI secondary structure................... 40

Figure 2.4: Kinetics of LUV disruption by 0.5 µM StAP PSI at 25º C ............................ 42

Figure 2.5: AFM height images of PE:PS bilayer patches at pH 4.5 on the native oxide

layer of a silicon wafer ................................................................................... 45

Figure 2.6: Tertiary structural features of StAP PSI ......................................................... 49

Figure 3.1: Crystal structure of StAP PSI (3RFI) ............................................................. 70

Figure 3.2: Sedimentation equilibrium analytical centrifugation analyses in iso-ionic

saline buffers at pH 3.0, 6.2 and 7.4............................................................... 76

Figure 3.3: Intrinsic Trp Fluorescence Emission .............................................................. 77

Figure 3.4: Far-UV CD spectra of StAP PSI .................................................................... 79

Figure 3.5: pH-dependence of PSI-induced LUV leakage ............................................... 81

Figure 3.6: Intrinsic Trp fluorescence emission of StAP PSI ........................................... 83

Figure 3.7: Comparison of timed far-UV CD spectra of PSI with 1:1:1

POPC:POPE:POPS LUVs (100 µM PL) in buffered saline pH 4.5 at 22° C 84

Figure 3.8: Cryo-Transmission Electron Microscopy images of 1:1 POPE:POPS LUVs in

buffered saline at pH 4.5 ................................................................................ 86

Figure 3.9: Summary of phenomena from cryo-TEM images .......................................... 88

Figure 3.10: PSI-treated vesicle size changes over time monitored by DLS .................... 90

ix

Figure 3.11: Leakage assays for peptide-treated 1:1 POPE:POPS LUVs at 25° C in

buffered saline pH 4.5 .................................................................................... 92

Figure 3.12: Intrinsic Trp fluorescence emission spectra upon incubation of 10 µM

peptide H3 with 1:1:1 POPC:POPE:POPS LUVs (100 µM PL) in buffered

saline pH 4.5 at 25° C .................................................................................... 94

Figure 3.13: Far-UV CD spectra of peptides in buffered saline containing 10 mM DTT at

the indicated pH values .................................................................................. 95

Figure 3.14: H3-treated vesicle size monitored over time by DLS .................................. 97

Figure 3.15: Protein domain structure alignments .......................................................... 110

Figure 3.16: Stability testing on LUV stocks after storage at ambient temperature ....... 123

Figure 4.1: PSI primary structure .................................................................................... 134

Figure 4.2: PSI-induced PL bilayer vesicle leakage in isoionic saline buffers at 25° C . 136

Figure 4.3: Bilayer fusion of 100 µM POPE:POPS as 100 nm LUVs ........................... 138

Figure 4.4: Bilayer fusion of 100 µM POPE:POPS as 100 nm LUVs in reducing

conditions ..................................................................................................... 139

Figure 4.5: Bilayer fusion of 100 µM LUVmix as 100 nm LUVs ................................. 141

Figure 4.6: Intrinsic Trp fluorescence emission spectra for StAP, phytepsin, cardosin A

and AtAP PSIs in buffered saline at varying pH.......................................... 144

Figure 4.7: Trp fluorescence emission of PSIs upon incubation with anionic bilayer

vesicles ......................................................................................................... 146

Figure 4.8: Far-UV CD spectra comparing PSIs at different pH in iso-ionic buffered

saline. Spectra are arranged for comparison between pH values for the

respective PSIs ............................................................................................. 153

Figure 4.9: Far-UV CD spectra comparing PSIs at different pH in iso-ionic buffered

saline. Spectra are arranged for comparison between PSIs at given pH values

...................................................................................................................... 154

Figure 5.1: Morphologies of PSI-treated LUVs observed by cryo-TEM ....................... 158

x

List of Abbreviations

AP Aspartic Protease

TRX Thioredoxin

PSI Plant Specific Insert

PSS Plant-Specific Sequence

SAPLIP Saposin-Like Protein

StAP Solanum tuberosum Aspartic Protease

DTT Dithiothreitol

LUV Large Unilamellar Vesicle

PC Phosphatidyl-choline

PE Phosphatidylethanolamine

PS Phosphatidylserine

AFM Atomic force microscopy

DLS Dynamic Light Scattering

RMSD Root mean square deviation

SNARE Soluble N-ethylmaleimide-sensitive factor attachment protein receptor

Trp Tryptophan

GRAS Generally Regarded as Safe

PL Phospholipid

POPC 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine

POPE 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoethanolamine

POPS 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoserine

CD Circular Dichroism Spectropolarimetry

TEM Cryo-Transmission Electron Microscopy

UV Ultraviolet

LUVmix Large unilamellar vesicles composed of a vacuole-like phospholipid mixture

1

Chapter 1: Introduction

1.1 Background

Proteases are an important class of enzymes due to their relevance in both the

physiological and commercial fields. Proteases are the most important type of industrial

use enzymes because they represent approximately 60% of all commercialized enzymes in

the world (Feijoo-Siota and Villa, 2011). The diverse fields of application include food

science and technology, pharmaceutical industries, and detergent manufacturing. Novo

Industries, Gist-Brocades, Genencor International, and Miles Laboratories are among the

top producers in the world (Feijoo-Siota and Villa, 2011). Aspartic proteases (APs)

constitute one of six classes of proteases (Seemuller, 1995). The studies herein consist of

protein structure-function investigations of the saposin-like domain of plant APs, also

known as the Plant Specific Insert (PSI) and the Plant Specific Sequence (PSS). This

introduction will focus on specific functional and structural details of plant AP PSIs,

however, the value of structure-function research in this area, as in the findings contained

in the proceeding chapters, extends beyond the scope of the chosen focus models: Potato

AP (StAP), Arabidopsis AP (AtAP), flowers of Cardoon AP (cardosin A) and barley AP

(phytepsin) (Devaraj, 2008).

Although the individual roles of different PSIs within their respective environments, e.g.

vacuoles, lysosomes, intercellular space, continue to be determined, there exists a clear

universal role for PSIs; to interact with membranes (Egas et al., 2000). Such interactions

can be with membranes of pathogenic invader species, e.g. potato leaf AP (Mendieta et al.,

2006), or membranes of endogenous cellular compartments in targeting roles, e.g. barley

2

AP (Tormakangas et al., 2001). Research in this area to date has focused on specific cases

involving PSIs of particular individual plants. There has been no apparent consensus on

the specific reasons that PSIs interact with membranes beyond the factual assertion that

PSIs are saposin-like protein structures. Lacking from the field of PSI structure-function

investigations has been the more global structure-function question: Are there structural

features common among PSIs that are universally responsible for their common function

as membrane-interacting domains on behalf of their AP protein carriers? To answer this

fundamental question, the present research will take a comprehensive approach that covers

SFRs of four PSIs, focusing on potato PSI. To accomplish this goal, differences in their

abilities to disrupt and/or fuse phospholipid bilayers, and structural comparisons at the

secondary, tertiary and quaternary protein structure levels as well as liposome morphology

will be exploited.

1.2 Literature Review

1.2.1 Plant aspartic proteinases - Throughout the plant kingdom, in angiosperms,

gymnosperms, and green algae, APs have been detected (Simoes and Faro, 2004). Plant

APs are distributed among various AP families; most plant APs are pepsin-like enzymes

belonging to the A1 AP family (Simoes and Faro, 2004). Plant APs are not as well

characterized as mammalian, microbial, and viral APs in terms of known structures and

biological roles (Mutlu and Gal, 1999, Simoes and Faro, 2004). APs have a bilobal

structure separated by an active site cleft, demonstrated well by the APs in in Figure 1.1

for distant members of the family (i.e., yeast vs. mammalian AP). Similar to pepsin, such

plant APs are most active and stable under acidic conditions having two catalytic Asp

3

Figure 1.1: The aspartic protease bilobal structure. The crystal structures of secreted

aspartic proteinase from Candida albicans (1ZAP) and human pepsin (1PSN) illustrate the

bilobal tertiary structure of aspartic proteases.

4

residues in the active site responsible for hydrolytic cleavage of protein substrate (Dunn,

2002; Simoes and Faro, 2004). Typically, plant AP N-termini have a pre-region containing

a signal peptide and a pro-region consisting of a peptide that binds to the active site usually

rendering the AP catalytically inactive as is found with most APs. In most known plant

APs the mature region contains an inserted sequence unique to plants called the plant-

specific sequence (PSS), also interchangeably called the plant-specific insert (PSI). The

term ‘insert’ is used to denote that this sequence is inserted into the C-terminal region of

mature sequences in plant zymogens (Figure 1.2). Usually PSIs are partially, or completely,

excised upon activation. The biological role(s) of this unique component of plant APs have

been studied since the first description of a unique plant AP component in 1991 (Runeberg-

Roos et al., 1991). The PSI is unlike any other AP component, but is highly similar to

saposin-like proteins NK-lysin (Liepinsh et al., 1997) (see Figures 1.3 and 1.4) and

granulysin (Anderson et al., 2003). Plant AP PSI primary structures are related to saposins

by a swapping of the N- and C-termini hence the term swaposin was coined to describe the

PSI (Ponting and Russell, 1995). Saposins are briefly described below in terms of saposin

structure-function relationships relevant to this proposal. PSI functions in plant APs are

still unclear; however, several roles have been suggested in the plant programmed cell

death (Faro et al., 1999, Tormakangas et al., 2001, Vieira et al., 2001). Specifically, the

presence of PSI in mature APs from potato (StAPs); tomato (LycoAP) (Schaller and Ryan,

1996) and Nepenthes alata (NaAP4; Philippine tropical pitcher plant) (An et al., 2002), has

been considered as being a part of the defensive machinery against pathogens and/or an

effector of cell death (An et al., 2002, Mendieta et al., 2006).

5

Figure 1.2: Plant aspartic protease insert. Schematic illustrating the position of the PSI

within the C-terminal domain of plant aspartic proteases. Pro- is the prosegment, DTG/DSG

are the catalytic consensus sequences, and heavy/light chain refers to the AP structure after post-

translational processing and removal of the PSI, a process that occurs in many but not all plant APs.

Figure 1.3: The crystal structure of prophytepsin (1QDM). Note the bilobal main

enzyme structure consisting of the N- and C-terminal domains, and the ~100 amino acid plant-

specific insert. The catalytic Asp residues (magenta) are located in a deep cleft between the N-

and C-terminal lobes, and the plant-specific insert (purple) is a separately folded domain whose

sequence occurs within the C-terminal half of the aspartic protease overall primary structure.

6

Figure 1.4: The canonical saposin fold. The three saposin-like proteins shown share a

common fold termed the “saposin fold”, first solved for NK-lysin (Liepinsh et al., 1997).

The structures shown are NK-lysin (1NKL), saposin C (2GTG) and phytepsin PSI

(1QDM).

7

1.2.2 Solanum tuberosum aspartic proteinases (StAP) - In studying the effect of abiotic

stress on potato tuber tissue protein content, including the nature of observed proteolytic

activities, led to the isolation and partial characterization of an aspartic proteinases (StAP

1) (Guevara et al., 1999). Proteolytic activity was assayed for potato tuber disk extracts

and an increase in activity of over 50% was observed after 24 h. Various inhibitors were

tested and only pepstatin strongly inhibited proteolytic activity, thus an AP was suspected.

The pepstatin inhibition was only prevalent for extracts of aerated tuber disks. Such

extracts were subjected to anion-exchange, and pepstatin-affinity, chromatography

yielding an approximately 40 kDa protein on reducing SDS-PAGE suggesting a monomer.

The resulting protein had a pH optimum of 5 and activity was dependent on the presence

of the reducing agent DTT (4mM). Glycosylation of potato tuber StAP was indicated by

binding to concanavalin A (Guevara et al., 1999).

To further elucidate the biological function(s) of StAP, another AP was isolated from

potato leaves (StAP 3) (Guevara et al., 2001). Extracts of wounded leaves were taken at 0

h, 24 h and 48 h. At 0 h and 24 h, 21% and 16% of proteolytic activity was inhibited by

pepstatin, respectively, compared to 62% inhibition after 48h indicating a four-fold

increase in the proportion of proteolytic activity due to APs and the induction of AP

expression. The pH optimum of leaf StAP was shown to be 3, lower than that of tuber

StAP, and activity was unaffected by reducing agent DTT. Also, leaf StAP did not bind to

concanavalin A indicating that it is not a glycoprotein (Guevara et al., 2001). The wound

response nature of AP expression in potato leaves was similar to the case of tomato leaf

AP induction and cauliflower seed AP induction (Fujikura and Karssen, 1995; Schaller and

Ryan, 1996), respectively, although APs are expressed in Arabidopsis constitutively

8

(Mutlu and Gal, 1999), and pathogenesis-related APs are also constitutively expressed in

tomato and tobacco (Rodrigo et al., 1991). Thus, wound or infection is not always

necessary for plant AP induction.

Subsequent to the above studies on StAPs, further characterization of tuber StAP

induction as it relates to pathogenic conditions was carried out for two field-tested potato

cultivars; one resistant and one susceptible to Phytophthora infestans, the cause of late

potato blight (Guevara et al., 2002). Interestingly, the resistant cultivar showed faster and

higher StAP induction post-wound infection in intercellular fluid. Additionally, the isolated

StAP showed in vitro inhibition of Phytophthora infestans and Fusarium solani. Although

the antimicrobial activity was absent for pure StAP samples inhibited by pepstatin,

intercellular washing fluid (IWF) displayed substantial antimicrobial activity in the

presence of pepstatin A. Apparently, the antimicrobial activity of StAP was at least in part

proteolytic in nature under the conditions tested, and StAP accounted for approximately

half of the antimicrobial activity of tuber IWF. However, neither pepsin nor trypsin

inhibited either fungus in antimicrobial proteolytic activity control reactions tested at

concentrations at the high end of the range used for StAP. These results suggested an

antimicrobial functionality for potato AP (Guevara et al., 2002).

The cDNA for potato leaf StAP was later isolated yielding a gene sequence that included

a PSI (Guevara et al., 2005a). Since isolated StAP from previous studies (Guevara et al.,

1999, Guevara et al., 2001) yielded monomers then mature, active StAP must retain its PSI

because the cDNA indicated that the putative PSI region is inserted into the C-terminal

domain of prepro-StAP (Guevara et al., 2005b). In terms of predicted amino acid sequence,

StAP PSI was shown to be similar to other plant PSIs with conservation of disulfide-bridge

9

Cys residues and saposin-like protein domain homologies consistent with StAP PSI being

a swaposin (Guevara et al., 2005b, Ponting and Russell, 1995). Guevara et al. (2005)

further tested for mRNA accumulation after P. infestans infection of one resistant and one

non-resistant cultivar, and found that the resistant cultivar produced higher levels of StAP

mRNA after infection further bolstering the role of StAP in the plant defense mechanism

against potato blight (Guevara et al., 2005b). Further evidence of such a biological role

was provided by showing direct membrane interaction for StAP with the fungal membranes

of F. solani and P. infestans as well as membrane permeabilization (Mendieta et al., 2006).

Both leaf and tuber StAPs displayed antifungal activities. Membrane binding was shown

by fluorescence microscopy of fluorescein isothiocyanate-labelled StAP and membrane

permeabilization was indicated by the uptake of a fluorescent dye by the test fungi

(Mendieta et al., 2006). These results suggest a possible role of the StAP PSI in the light

of saposin proteins’ functions of membrane interaction and permeabilization (outlined

below). StAPs also display similar activities against bovine and human permatozoa in that

motility was abolished and membrane permeabilization was shown by the same

fluorescence-based method discussed above (Cesari et al., 2007).

Solanum tuberosum aspartic protease (StAP) PSI is able to kill spores of two potato

pathogens in a dose-dependent manner without any deleterious effect on plant cells (Muñoz

et al., 2010). The StAP-PSI ability to kill microbial pathogens is dependent on the direct

interaction of the protein with the microbial cell wall/or membrane, leading to increased

permeability and lysis. StAP-PSI and StAPs are cytotoxic to Gram-negative and Gram-

positive bacteria in a dose dependent manner, and furthermore, StAP-PSI is able to kill

human pathogenic bacteria in a dose dependent manner, but is not toxic to human red blood

10

cells at the concentrations and times assayed. MBC values determined for StAPs and StAP-

PSI are in the same order of magnitude as those previously reported for NK-lysin and

granulysin (Muñoz et al., 2010). Of interest is the structural basis for this selective toxicity

among StAP-PSI and other SAPLIPs. More recently, Muñoz et al. (2011) proposed that

the possible role of StAPs as pathogenesis related proteins into the plant defense response

is to interact with the pathogen plasma membrane causing membrane destabilization and

subsequent cell death and/or kill plant cells under stress conditions, thereby attenuating

pathogen spread and colonization (Muñoz et al., 2011).

The structure-function role of StAP glycosylation has also been studied (Pagano et al.,

2006, Pagano et al., 2007). De-glycosylation of StAPs did not alter pH-optimum for

activity, temperature optimum for activity, nor index of surface hydrophobicity (Pagano et

al., 2007). However, secretion into the intercellular space did not occur in response to

wounding and infection as would normally be expected for StAPs. Additionally, in vitro

antifungal activity was reduced substantially at various protein concentrations. It was thus

concluded that StAP glycosylation was essential for in vivo plant defense, but structure and

proteolytic activity was not affected (Pagano et al., 2007).

1.2.3 Saposins - The PSIs of APs are saposin-like proteins. Reviews of saposin-like

proteins were most recently published in 2005 (Bruhn, 2005b; Kolter and Sandhoff, 2005).

Saposins A, B, C, and D belong to the saposin-like protein family of proteins. Proteins of

this family are membrane-interacting in three principal ways: Membrane binding,

membrane perturbation without permeabilization, and membrane permeabilization (Bruhn,

2005b). The saposins are proteins that enhance the activities of lysosomal exohydrolases

in glycosphingolipid degradation (Matsuda et al., 2001). Saposins A-C are implicated in

11

various disease states whereas no known deficiency of saposin D in humans has been

documented, however, a saposin D mouse knockout resulted in deleterious effects

(Matsuda, 2008). In general, defective saposin-disease states arise from the accumulation

of ceramide derivatives in various tissues resulting in pathological states. Saposins are

relatively small at 8 - 11 kD, they are heat stable, and they display non-enzymatic activities

(Kolter and Sandhoff, 2005). Also, they contain three disulfide bonds holding together a

five-helix, compact, globular structure that has been termed a ‘saposin-fold structure’ (You

et al., 2003). The four saposins are all derived from the single precursor protein prosaposin,

a protein sequence made up of the four saposins in tandem which are subsequently

proteolytically processed in the lysosome thereby yielding the four individual products

(Matsuda, 2008).

Atomic force microscopy revealed the effects of saposin C interactions on phospholipid-

containing membranes (You et al., 2003). Two main effects were visualized: membrane

regions were rearranged to form so-called patch-like domains (literally irregularly-shaped,

circular regions of increased thickness) and membrane destabilization. Unlike the

formation of patch-like domains, membrane destabilization was dependent on the presence

and concentration of acidic phospholipids, e.g., phosphatidylserine (You et al., 2003). Of

particular interest in this study, the first two helices of saposin C were synthesized and

tested independently. Helix 1 was found to be incapable of membrane restructuring alone,

however it destabilized membrane if it was added first followed by helix 2. By contrast,

helix 2 was capable of patch-like domain formation only (You et al., 2003). Individual

structural regions of saposins within the saposin fold are thus proven to play specific roles

with respect to membrane interactions.

12

Study of saposin structure by X-ray crystallography suggested that human saposin B

bound to phosphatidylethanolamine exists as a dimer (PDB 1N69) (Ahn et al., 2003b). The

secondary structure of bound dimer was similar to that of the monomer, however, dimer

contained rearranged helices at the tertiary structure level. Overall, a relatively large

hydrophobic cavity (relative to the small saposin structure) is enclosed by a monolayer of

helices. The large hydrophobic cavity could serve as an area of direct interaction with

membrane surfaces (Ahn et al., 2003b). This “open” saposin fold was later observed again

for saposin C (PDB 2QYP) such that two extended monomer form a homodimer. The

closed and open saposin folds are shown for the same protein, saposin C, in Figure 1.5.

The existence of the open saposin fold was particularly important to understanding the

mechanism by which saposin C induces bilayer fusion (Rossmann et al., 2008) in that it

gave credence to the previously proposed “clip-on” mechanism. In this model, each of the

saposin C monomer units bind adjacent bilayers thereby bringing them into proximity, and

due to their having perturbed the bilayer structures, adjacent damaged lipid structure will

have an increased likelihood of joining upon rearrangement of an energetically favourable

state.

1.2.4 Structure-Function Studies of Saposin Disulfide Bonds - The saposin fold has

been shown to interact with membranes (You et al., 2003). All four saposins contain six

identically placed, conserved cysteine residues (Vaccaro et al., 1994). The highly

conserved nature of disulfide bonds suggests an important biological role for the bonds’

effects on saposin structures. Establishing the location of saposin disulfide bonds was

considered to be critical in assessing saposin conformations and their importance for

saposin-lipid bilayer interactions (Vaccaro et al., 1995).

13

Figure 1.5: Comparison of the two known saposin folds. Top: The ‘closed” saposin fold

of saposin C (2GTG). This is the canonical fold for the saposin-like family. Bottom: The “open”

saposin fold of saposin C (2QYP). The open fold is the only known fold for saposin dimerization.

14

1.2.5 Plant AP PSI-Membrane Interactions - Proteins belonging to the Saposin-Like

Protein Family have various physiological functions that have been suggested: Saposins

are critical to exohydrolase degradation of sphingolipids in the lysosome, NK-lysin is

antibacterial and tumor-lysing, granulysin is antimicrobial, surfactant protein B lowers

pulmonary surfactant surface tension, and amoebapores lyse bacteria and eukaryotic cells.

Membrane interaction is the common theme among all of these functions (Egas et al.,

2000). Beyond the direct study of saposin-like proteins’ primary functions, the

understanding of structure-function relationships involving membrane-interacting AP

surfaces could also extend to the membrane-bound APs, i.e. memapsins 1 and 2, implicated

in Alzheimer’s disease, which do not contain PSIs. The interaction of procardosin A with

membrane vesicles has been studied (Egas et al., 2000). The role of the PSI for membrane

interactions of procardosin A with phospholipid vesicles was explored by recombinantly

expressing procardosin A, a PSI-deletion mutant procardosin A, and PSI in E. coli. Full

length procardosin caused membrane damage as evidenced by vesicle leakage whereas

PSI-deletion mutant procardosin A was ineffective at disrupting lipid bilayers (Egas et al.,

2000). The PSI and full-length enzyme leakage results suggested that the PSI is responsible

for membrane interactions and hence may function as part of both a host defense

mechanism as well as programmed cell death.

1.3 Scientific Investigation Plan

1.3.1 Research Questions - The present studies are not only important to improving the

current state of knowledge and understanding of PSI structure-function relationships, but

are prerequisite to future engineering of PSIs or PSI-templated interventions for disease

resistance, e.g., potato blight resistance, or disease treatments including human medical

15

interventions. In conceptualizing an approach towards the above, several questions related

to plant AP PSIs were formulated:

(1) Are plant PSI secondary and tertiary structures self-determined from their primary

sequences?

(2) What structural characteristics determine observed bilayer interactions?

(3) How do swaposins (PSIs) compare with saposins in terms of bilayer effects?

(4) What is the mechanism by which PSI disrupts bilayers?

(5) Are there important functional differences between the PSIs of four representative plant

species; potato, barley, flowers of Cardoon (thistle) and Arabidopsis? That is - is the

primary sequence of a swaposin a universal determinant of bilayer disruption/interaction?

1.3.2 Objectives -The overarching objective of the present research projects are as

follows:

1. To delineate the structural basis of potato PSI-PL bilayer interactions with respect to

protein folding and structure, and bilayer PL composition.

2. To assess the universality of PSI structure-function relationships by comparing PSIs of

four different plant species in terms of their bilayer interactions and protein structural

features.

16

1.3.3 Hypotheses

1. The PSI (swaposin) domain of potato AP is an independently functional protein unit,

independent of its parent AP for its bilayer disruption activity, and this bilayer

perturbation occurs in a similar manner to saposins.

2. The quaternary structure of StAP PSI, i.e., dimerization, is critical to catalyzing bilayer

fusion whereas bilayer disruption activity is dependent on the C-terminal portion of the

PSI molecule, independent of tertiary or quaternary structure.

3. The bilayer effects of the PSIs of potato, barley, flowers of Cardoon and Arabidopsis

are indistinguishable in terms of membrane disruption and bilayer fusion.

17

1.4 References

Ahn, V.E., Faull, K.F., Whitelegge, J.P., Fluharty, A.L. & Privé, G.G. 2003. Crystal

Structure of Saposin B Reveals a Dimeric Shell for Lipid Binding. Proceedings of the

National Academy of Sciences of the United States of America, 100(1): 38-43.

Alattia, J.R., Shaw, J.E., Yip, C.M. & Prive, G.G. 2006. Direct Visualization of Saposin

Remodeling of Lipid Bilayers. Journal of Molecular Biology, 362(5): 943-953.

An, C., Fukusaki, E. & Kobayashi, A. 2002. Aspartic Proteinases Are Expressed in Pitchers

of the Carnivorous Plant Nepenthes Alata Blanco. Planta, 214(5): 661-667.

Anderson, D.H., Sawaya, M.R., Cascio, D., Ernst, W., Modlin, R., Krensky, A. &

Eisenberg, D. 2003. Granulysin Crystal Structure and a Structure-Derived Lytic

Mechanism. Journal of Molecular Biology, 325(2): 355-365.

Baker, D. & Agard, D.A. Influenza Hemagglutinin: Kinetic Control of Protein Function.

Cell, 2(10): 907-910.

Bartlett, G.R. 1959. Phosphorous Assay in Column Chromatography. Journal of Biological

Chemistry, 234(3): 466-468.

Braun, H. & Schmitz, U.K. 1995. Are the ‘core’ Proteins of the Mitochondrial bc1

Complex Evolutionary Relics of a Processing Protease? Trends in Biochemical

Sciences, 20(5): 171-175.

Braun, H. & Schmitz, U.K. 1999. The Protein-Import Apparatus of Plant Mitochondria.

Planta, 209(3): 267-274.

Bruhn, H. 2005. A Short Guided Tour through Functional and Structural Features of

Saposin-Like Proteins. Biochemical Journal, 389: 249-257.

Cesari, A., Falcinelli, A.L., Mendieta, J.R., Pagano, M.R., Mucci, N., Daleo, G.R. &

Guevara, M.G. 2007. Potato Aspartic Proteases (StAPs) Exert Cytotoxic Activity on

Bovine and Human Spermatozoa. Fertility and sterility, 88(4, Supplement 1): 1248-

1255.

Chernomordik, L.V., Frolov, V.A., Leikina, E., Bronk, P. & Zimmerberg, J. 1998. The

Pathway of Membrane Fusion Catalyzed by Influenza Hemagglutinin: Restriction of

Lipids, Hemifusion, and Lipidic Fusion Pore Formation. The Journal of cell biology,

140(6): 1369-1382.

18

Ciaffoni, F., Salvioli, R., Tatti, M., Arancia, G., Crateri, P. & Vaccaro, A.M. 2001. Saposin

D Solubilizes Anionic Phospholipid-Containing Membranes. Journal of Biological

Chemistry, 276(34): 31583-31589.

Ciaffoni, F., Tatti, M., Boe, A., Salvioli, R., Fluharty, A., Sonnino, S. & Vaccaro, A.M.

2006. Saposin B Binds and Transfers Phospholipids. Journal of lipid research, 47(5):

1045-1053.

Copley, S.D. 2003. Enzymes with Extra Talents: Moonlighting Functions and Catalytic

Promiscuity. Current opinion in chemical biology, 7(2): 265-272.

Devaraj, K.B., Gowda, L.R., Prakash, V. 2008. An Unusual Thermostable Aspartic

Protease from the Latex of Ficus Racemosa (L.). Phytochemistry, 69(3): 647-655.

Dunn, B.M. 2002. Structure and Mechanism of the Pepsin-Like Family of Aspartic

Peptidases. Chemical reviews, 102(12): 4431-4458.

Egas, C., Lavoura, N., Resende, R., Brito, R.M., Pires, E., de Lima, M.C. & Faro, C. 2000.

The Saposin-Like Domain of the Plant Aspartic Proteinase Precursor is a Potent Inducer

of Vesicle Leakage. The Journal of biological chemistry, 275(49): 38190-38196.

Egawa, H. & Furusawa, K. 1999. Liposome Adhesion on Mica Surface Studied by Atomic

Force Microscopy. Langmuir, 15(5): 1660-1666.

Faro, C., Ramalho-Santos, M., Vieira, M., Mendes, A., Simoes, I., Andrade, R., Verissimo,

P., Lin, X., Tang, J. & Pires, E. 1999. Cloning and Characterization of cDNA Encoding

Cardosin A, an RGD-Containing Plant Aspartic Proteinase. J.Biol.Chem., 274(40):

28724-28729.

Feijoo-Siota, L. & Villa, T. 2011. Native and Biotechnologically Engineered Plant

Proteases with Industrial Applications. Food and Bioprocess Technology, 4(6): 1066-

1088.

Fiske, C.H. & Subbarow, Y. 1925. The Colorimetric Determination of Phosphorous. The

Journal of Biological Chemistry, 66(2): 375-400.

Fujikura, Y. & Karssen, C.M. 1995. Molecular Studies on Osmoprimed Seeds of

Cauliflower: A Partial Amino Acid Sequence of a Vigour-Related Protein and

Osmopriming-Enhanced Expression of Putative Aspartic Protease. Seed Science

Research, 5(3): 177-181.

19

Guevara, M.G., Oliva, C.R., Machinandiarena, M. & Daleo, G.R. 1999. Purification and

Properties of an Aspartic Protease from Potato Tuber that is Inhibited by a Basic

Chitinase. Physiologia Plantarum, 106(2): 164-169.

Guevara, M.G., Almeida, C., Mendieta, J.R., Faro, C.J., Verissimo, P., Pires, E.V. & Daleo,

G.R. 2005. Molecular Cloning of a Potato Leaf cDNA Encoding an Aspartic Protease

(StAsp) and its Expression After P. Infestans Infection. Plant physiology and

Biochemistry, 43(9): 882-889.

Guevara, M.G., Daleo, G.R. & Oliva, C.R. 2001. Purification and Characterization of an

Aspartic Protease from Potato Leaves. Physiologia plantarum, 112(3): 321-326.

Guevara, M.G., Oliva, C.R., Huarte, M. & Daleo, G.R. 2002. An Aspartic Protease with

Antimicrobial Activity is Induced After Infection and Wounding in Intercellular Fluids

of Potato Tubers. European Journal of Plant Pathology, 108(2): 131-137.

Han, X., Bushweller, J.H., Cafiso, D.S. & Tamm, L.K. 2001. Membrane Structure and

Fusion-Triggering Conformational Change of the Fusion Domain from Influenza

Hemagglutinin. Nature Structural Biology, 8: 715-720.

Han, X. & Tamm, L.K. 2000. A host–guest System to Study structure–function

Relationships of Membrane Fusion Peptides. Proceedings of the National Academy of

Sciences of the United States of America, 97(24): 13097-13102.

Hernandez, L.D., Hoffman, L.R., Wolfsberg, T.G. & White, J.M. 1996. Virus-Cell and

Cell-Cell Fusion. Annual Review of Cell and Developmental Biology, 12: 627-627-661.

Huster, D., Arnold, K. & Gawrisch, K. 2000. Strength of Ca2+ Binding to Retinal Lipid

Membranes: Consequences for Lipid Organization. Biophysical journal, 78(6): 3011-

3018.

Jeffery, C.J. 2009. Moonlighting Proteins - An Update. Molecular BioSystems, 5(4): 345-

350.

Kervinen, J., Tobin, G.J., Costa, J., Waugh, D.S., Wlodawer, A. & Zdanov, A. 1999.

Crystal Structure of Plant Aspartic Proteinase Prophytepsin: Inactivation and Vacuolar

Targeting. EMBO Journal, 18(14): 3947-3955.

Kingsley, D.H., Behbahani, A., Rashtian, A., Blissard, G.W. & Zimmerberg, J. 1999. A

Discrete Stage of Baculovirus GP64-Mediated Membrane Fusion. Molecular Biology

of the Cell, 10(12): 4191-4200.

20

Kolter, T. & Sandhoff, K. 2005. Principles of Lysosomal Membrane Digestion:

Stimulation of Sphingolipid Degradation by Sphingolipid Activator Proteins and

Anionic Lysosomal Lipids. Annual Review of Cell and Developmental Biology, 21(1):

81-103.

Laemmli, U.K. 1970. Cleavage of Structural Proteins during the Assembly of the Head of

Bacteriophage T4. Nature, 227: 680-685.

Lai, A.L., Park, H., White, J.M. & Tamm, L.K. 2006. Fusion Peptide of Influenza

Hemagglutinin Requires a Fixed Angle Boomerang Structure for Activity. Journal of

Biological Chemistry, 281(9): 5760-5770.

Lai, A.L. & Tamm, L.K. 2010. Shallow Boomerang-Shaped Influenza Hemagglutinin

G13A Mutant Structure Promotes Leaky Membrane Fusion. Journal of Biological

Chemistry, 285(48): 37467-37475.

Liepinsh, E., Andersson, M., Ruysschaert, J.M. & Otting, G. 1997. Saposin Fold Revealed

by the NMR Structure of NK-Lysin. Nature Structural Biology, 4(10): 793-795.

MacDonald, R.C., MacDonald, R.I., Menco, B.P.M., Takeshita, K., Subbarao, N.K. & Hu,

L. 1991. Small-Volume Extrusion Apparatus for Preparation of Large, Unilamellar

Vesicles. Biochimica et Biophysica Acta - Biomembranes, 1061(2): 297-303.

Markovic, I., Pulyaeva, H., Sokoloff, A. & Chernomordik, L.V. 1998. Membrane Fusion

Mediated by Baculovirus gp64 Involves Assembly of Stable gp64 Trimers into

Multiprotein Aggregates. Journal of Cell Biology, 143(5): 1155-1166.

Matsuda, J. 2008. Sphingolipid Activator Proteins, In Experimental Glycoscience, Eds.

Taniguchi, N., Suzuki, A., Ito, Y., Narimatsu, H., Kawasaki, T. and Hase, S., Springer

Japan, Tokyo, pp. 125-129.

Matsuda, J., Vanier, M.T., Saito, Y., Tohyama, J., Suzuki, K. & Suzuki, K. 2001. A

Mutation in the Saposin A Domain of the Sphingolipid Activator Protein (Prosaposin)

Gene Results in a Late-Onset, Chronic Form of Globoid Cell Leukodystrophy in the

Mouse. Human molecular genetics, 10(11): 1191-1199.

Matsuzaki, K., Harada, M., Handa, T., Funakoshi, S., Fujii, N., Yajima, H. & Miyajima,

K. 1989. Magainin 1-Induced Leakage of Entrapped Calcein Out of Negatively-Charged

Lipid Vesicles. Biochimica et Biophysica Acta - Biomembranes, 981(1): 130-134.

21

Mendieta, J.R., Pagano, M.R., Munoz, F.F., Daleo, G.R. & Guevara, M.G. 2006.

Antimicrobial Activity of Potato Aspartic Proteases (StAPs) Involves Membrane

Permeabilization. Microbiology, 152: 2039-2047.

Moore, B.d. 2004. Bifunctional and Moonlighting Enzymes: Lighting the Way to

Regulatory Control. Trends in Plant Science, 9(5): 221-228.

Muñoz, F.F., Mendieta, J.R., Pagano, M.R., Paggi, R.A., Daleo, G.R. & Guevara, M.G.

2010. The Swaposin-Like Domain of Potato Aspartic Protease (StAsp-PSI) Exerts

Antimicrobial Activity on Plant and Human Pathogens. Peptides, 31(5): 777-785.

Muñoz, F., Palomares-Jerez, M.F., Daleo, G., Villalaín, J. & Guevara, M.G. 2011.

Cholesterol and Membrane Phospholipid Compositions Modulate the Leakage Capacity

of the Swaposin Domain from a Potato Aspartic Protease (StAsp-PSI). Biochimica et

Biophysica Acta - Molecular and Cell Biology of Lipids, 1811(12): 1038-1044.

Mutlu, A. & Gal, S. 1999. Plant Aspartic Proteinases: Enzymes on the Way to a Function.

Physiologia Plantarum, 105(3): 569-576.

Pagano, M.R., Mendieta, J.R., Munoz, F.F., Daleo, G.R. & Guevara, M.G. 2006. Role of

Glycosylation on Potato Aspartic Proteases Secretion. Phytopathology, 96(6): S88-S89.

Pagano, M.R., Mendieta, J.R., Munoz, F.F., Daleo, G.R. & Guevara, M.G. 2007. Roles of

Glycosylation on the Antifungal Activity and Apoplast Accumulation of StAPs

(Solanum Tuberosum Aspartic Proteases). International Journal of Biological

Macromolecules, 41(5): 512-520.

Paumet, F., Rahimian, V., Di Liberto, M. & Rothman, J.E. 2005. Concerted Auto-

Regulation in Yeast Endosomal t-SNAREs. Journal of Biological Chemistry, 280(22):

21137-21143.

Paumet, F., Rahimian, V. & Rothman, J.E. 2004. The Specificity of SNARE-Dependent

Fusion is Encoded in the SNARE Motif. Proceedings of the National Academy of

Sciences of the United States of America, 101(10): 3376-3380.

Petrache, H.I., Tristram-Nagle, S. & Nagle, J.F. 1998. Fluid Phase Structure of EPC and

DMPC Bilayers. Chemistry and Physics of Lipids, 95(1): 83-94.

Ponting, C.P. & Russell, R.B. 1995. Swaposins: Circular Permutations within Genes

Encoding Saposin Homologues. Trends in Biochemical Sciences, 20(5): 179-180.

22

Remmel, N., Locatelli-Hoops, S., Breiden, B., Schwarzmann, G. & Sandhoff, K. 2007.

Saposin B Mobilizes Lipids from Cholesterol-Poor and Bis-(Monoacylglycero)-

Phosphate-Rich Membranes at Acidic pH. FEBS J, 274(13): 3405-3420.

Rodrigo, I., Vera, P., Van Loon, L.C. & Conejero, V. 1991. Degradation of Tobacco

Pathogenesis-Related Proteins: Evidence for Conserved Mechanisms of Degradation of

Pathogenesis-Related Proteins in Plants. Plant Physiology, 95(2): 616-622.

Rossmann M., Schultz-Heienbrok R., Behlke J., Remmel N., Alings C., Sandhoff K.,

Saenger W. and Maier T. (2008) Crystal Structures of Human Saposins C and D:

Implications for Lipid Recognition and Membrane Interactions. Structure, 16: 809-817.

Runeberg-Roos, P., Tormakangas, K. & Ostman, A. 1991. Primary Structure of a Barley-

Grain Aspartic Proteinase. A Plant Aspartic Proteinase Resembling Mammalian

Cathepsin D. European Journal of Biochemistry, 202(3): 1021-1027.

Schaller, A. & Ryan, C.A. 1996. Molecular Cloning of a Tomato Leaf cDNA Encoding an

Aspartic Protease, a Systemic Wound Response Protein. Plant Molecular Biology,

31(5): 1073-1077.

Sharom, F.J., DiDiodato, G., Yu, X. & Ashbourne, K.J.D. 1995. Interaction of the P-

Glycoprotein Multidrug Transporter with Peptides and Ionophores. Journal of

Biological Chemistry, 270(17): 10334-10341.

Seemuller, E., Lupas, A., Stock, D., & Lowe, J. 1995. Proteasome from Thermoplasma

acidophilum: a threonine protease. Science 268(5210): 579-582.

Simoes, I. & Faro, C. 2004. Structure and Function of Plant Aspartic Proteinases. The FEBS

Journal, 271(11): 2067-2075.

Sreerama, N. & Woody, R.W. 2004. Computation and Analysis of Protein Circular

Dichroism Spectra. Methods in Enzymology, 383: 318-345.

Tormakangas, K., Hadlington, J.L., Pimpl, P., Hillmer, S., Brandizzi, F., Teeri, T.H. &

Denecke, J. 2001. A Vacuolar Sorting Domain may also Influence the Way in which

Proteins Leave the Endoplasmic Reticulum. The Plant Cell, 13(9): 2021-2032.

Vaccaro, A.M., Tatti, M., Ciaffoni, F., Salvioli, R., Serafino, A. & Barca, A. 1994. Saposin

C Induces pH-Dependent Destabilization and Fusion of Phosphatidylserine-Containing

Vesicles. FEBS Letters, 349(2): 181-186.

23

Vaccaro, A.M., Salvioli, R., Barca, A., Tatti, M., Ciaffoni, F., Maras, B., Siciliano, R.,

Zappacosta, F., Amoresano, A. & Pucci, P. 1995. Structural Analysis of Saposin C and

B. Journal of Biological Chemistry, 270(17): 9953-9960.

Vieira, M., Pissarra, J., Veríssimo, P., Castanheira, P., Costa, Y., Pires, E. & Faro, C. 2001.

Molecular Cloning and Characterization of cDNA Encoding Cardosin B, an Aspartic

Proteinase Accumulating Extracellularly in the Transmitting Tissue of Cynara

Cardunculus L. Plant Molecular Biology, 45(5): 529-539.

Wang, Y., Grabowski, G.A. & Qi, X. 2003. Phospholipid Vesicle Fusion Induced by

Saposin C. Archives of Biochemistry and Biophysics, 415(1): 43-53.

You, H.X., Qi, X., Grabowski, G.A. & Yu, L. 2003. Phospholipid Membrane Interactions

of Saposin C: In Situ Atomic Force Microscopic Study. Biophysical Journal, 84(3):

2043-2057.

24

Chapter 2: Structure and Mechanism of the Saposin-Like

Domain of a Plant Aspartic Protease

Note: The content of this chapter was published in The Journal of Biological Chemistry

(Bryksa et al., 2011). The article includes reporting the crystal structure of StAP PSI

(section 2.4.1), work done by Dr. Prasenjit Bhaumik, Eugenia Magracheva, Dr. Alexander

Zdanov and Dr. Alexander Wlodawer (SAIC-Frederick, National Cancer Institute,

Frederick, MD 21702, USA).

2.1 Abstract

Many plant aspartic proteases contain an additional sequence of approximately 100

amino acids termed the plant-specific insert which is involved in host defense and vacuolar

targeting. Similar to all saposin-like proteins, the plant-specific insert functions via protein-

membrane interactions, however, the structural basis for such interactions has not been

studied and the nature of plant-specific insert mediated membrane disruption has not been

characterized. In the present study, the crystal structure of the saposin-like domain of potato

aspartic protease was resolved at a resolution of 1.9 Å, revealing an open V-shaped

configuration similar to the open structure of human saposin C. Notably vesicle disruption

activity followed Michaelis-Menten-like kinetics, a finding not previously reported for

saposin-like proteins including plant-specific inserts. Circular dichroism data suggested

that secondary structure was pH-dependent in a fashion similar to influenza A

hemagglutinin fusion peptide. Membrane effects characterized by atomic force microscopy

and light scattering indicated bilayer solubilization as well as fusogenic activity. Taken

together, the present study is the first report to elucidate the membrane interaction

mechanism of plant saposin-like domains whereby pH-dependent membrane interactions

25

resulted in bilayer fusogenic activity that likely arose from a viral-type pH-dependent helix-

kink-helix motif at the plant-specific insert N-terminus.

2.2 Introduction

Aspartic proteases (APs) are characterized by a common bilobal tertiary structure

containing two catalytic aspartic acid residues (Asp32 and Asp215 in pepsin) within an

active site cleft (Blundell & Johnson, 1993; Davies, 1990). They are found in all higher

organisms and their respective roles are well established, although structural and functional

characteristics of APs in plants are least understood. Of practical interest among plant APs

are their roles in plant pathogen resistance (Guevara et al, 2002) as well as in senescence

and post-harvest physiology (Payie et al, 2000; Schaller & Ryan, 1996). Plant APs share

the common AP bilobal structure; however, some contain an additional sequence of

approximately 100 residues inserted within the C-terminal primary structure. These

additional amino acids unique to plant APs (Glathe et al, 1998; Payie et al, 2003; Ramalho-

Santos et al, 1998) create an extra domain protruding from the canonical AP molecule

(Frazão et al, 1999; Kervinen et al, 1999; Mazorra-Manzano et al, 2010). This structural

oddity among APs is called the plant-specific insert (PSI), also known as the plant-specific

sequence (PSS), which belongs to the saposin-like protein (SAPLIP) family (Mutlu & Gal,

1999; Runeberg‐Roos et al, 1991). Plant APs are found in either monomeric or

heterodimeric forms (Egas et al, 2000; Kervinen et al, 1999); the latter result from post-

26

translational proteolysis which includes the removal of part or all of the PSI, whereas the

PSI is retained in monomeric plant APs (Glathe et al, 1998; Ramalho-Santos et al, 1998).

In general, members of the SAPLIP family have various physiological functions all of

which entail membrane interaction (Bruhn, 2005; Egas et al, 2000; Kolter & Sandhoff,

2005) manifested in three principal ways: membrane binding, membrane perturbation

without permeabilization, and membrane permeabilization (Bruhn, 2005). Examples of

SAPLIP functions include roles in exohydrolase degradation of sphingolipids in the

lysosome (saposins) (Matsuda et al, 2001), antimicrobial activity (granulysin and NK-

lysin) (Anderson et al, 2003), tumor lysis (NK-lysin) (Liepinsh et al, 1997), pulmonary

surfactant surface tension regulation (surfactant protein B) (Gordon et al, 2000), and

bacterial/eukaryotic cell lysis (amoebapores) (Zhai & Saier, 2000).

Fusion of cellular lipid membranes is an essential process in all forms of life (Stiasny et

al, 2007) and the mechanism by which membrane fusion occurs, a process typically

catalyzed by proteins, continues to be unraveled (Kasson et al, 2010). Merely bringing

membranes in proximity to one another is insufficient for fusion (Kasson et al, 2010), and

the nature of fusion peptide structures is critical to fusogenic function (Bissonnette et al,

2009; Lai et al, 2006; Qiao et al, 1999). Disordering of bilayers by fusion proteins, thought

to be a critical first step in the catalysis of bilayer fusion (Lai et al, 2006), results in an

increased rate of energetically unfavorable hydrophobic lipid tail protrusion (Kasson et al,

2010). The fusion transition state involves contact formation between lipid tails of opposite

bilayers within the intervening hydrophilic region (Kasson et al, 2010) resulting in stalk

formation(s) between the two disordered bilayer patches (Kasson et al, 2010; Lukatsky &

Frenkel, 2004). Dimerization of helical structures is part of the saposin-mediated bilayer

27

fusion, transfer, and solubilizing mechanisms (Ciaffoni et al, 2006; Remmel et al, 2007;

Wang et al, 2003) and these structural rearrangements take place after release from the

parent molecule (prosaposin) (Kolter & Sandhoff, 2005).

The SAPLIP domains of plant APs display membrane permeabilizing activity

independent of its “parent” protein (Egas et al, 2000; Simoes & Faro, 2004) and they likely

act independently (post-proteolytic processing) as a part of the plant defense mechanism

against fungal pathogens (Guevara et al, 2002; Muñoz et al, 2010). Like PSIs of

heterodimeric plant APs, saposins are also expressed as a proprotein and are subsequently

processed via proteolytic cleavage (Bruhn, 2005) resulting in distinct, active tertiary

structures consisting of stable helical and coil secondary structures (Ahn et al, 2003; Ahn

et al, 2006; Alattia et al, 2006; Bruhn, 2005; Ciaffoni et al, 2006).

Recently, recombinantly-produced PSI of Solanum tuberosum aspartic proteinase

(StAP) was shown to kill human pathogens as well as inhibit fungal sporulation via

interaction with, and permeabilization of, microbial plasma membranes (Muñoz et al,

2010). Understanding the structural basis for newly characterized antifungal activities is

important in the development of novel therapeutic drugs for the treatment of fungal

infections (Ghannoum & Rice, 1999) in immunocompromised patients (Meyer, 2008;

Tseng & Perfect, 2011). Furthermore, we propose that understanding structure-function

relationships involving PSI-membrane interactions may have relevance to non-plant

membrane-bound APs, e.g. memapsins 1 and 2, implicated in Alzheimer’s disease, beyond

the direct elucidation of SAPLIP primary functions. Using StAP PSI as a model system,

the present study characterizes the structure of a plant AP PSI as it relates to membrane

interactions. The observed saposin C-like tertiary structure and saposin B-like fusogenic

28

activity, and the apparent catalysis of energetically unfavorable membrane bilayer

disruption and fusion via a pH-dependent helix fusion peptide motif at the PSI N-terminus

are discussed.

2.3 Experimental Procedures

2.3.1 Materials - A PSI synthetic gene optimized for expression in E. coli was purchased

from Mr. Gene GmbH (Regensburg, Germany). Plasmids pET19b(+) and pET32b(+), E.

coli Rosetta-gami B (DE3)pLysS, and u-MACTM columns were obtained from EMD

Biosciences (San Diego, CA, USA). E. coli TOP10F’ was from Invitrogen (San Diego,

CA, USA). GenEluteTM Plasmid Miniprep Kit was obtained from Sigma-Aldrich Co. (St.

Louis, MO, USA). The QIAquick® PCR Purification Kit and QIAquick® Gel Extraction

Kit were from Qiagen (Germantown, MD, USA). Restriction enzymes, T4 DNA ligase and

Pfu DNA polymerase were obtained from Fermentas Life Sciences (Burlington, ON,

Canada). Primers were synthesized by Sigma Genosys (Oakville, ON, Canada) and

thrombin was purchased from Fisher Scientific Co. (Ottawa, ON, Canada). The RPC

column was from GE Healthcare (Piscataway, NJ, USA). Phospholipids were from Avanti

Polar Lipids (Alabaster, AL, USA).

2.3.2 Construction - of Expression Vector pET32z-PSI: Two constructs, a 6.0 kb

construct named pET19b-PSI and a 6.2 kb construct named pET32z-PSI, were made for

the expression of PSI and thioredoxin-PSI fusion protein (Trx-PSI), respectively, in E. coli.

For pET19b-PSI, PSI insert was amplified using primers:

FwdPSINdeI 5’CATATGATTGTAAGCATGGAGTGTAAAACC and

RevPSIXhoI 5’ATCTCGAGTTACGGGATTTTTTCACACAGTTG,

29

followed by ligation between the NdeI and XhoI restriction sites. pET32z-PSI construct was

made using PSI insert amplified using primers:

FwdPSINcoI 5’ATCCATGGCGATTGTAAGCATGGAGTGTAAAACC and

RevPSIXhoI, followed by ligation between the NcoI and XhoI restriction sites of pET32z,

a modified version of pET32b which contains a deletion between the thrombin cut site and

the end of the enterokinase cut site. Each construct was transformed into E. coli TOP10F’

using the method of Hanahan (Hanahan, 1983).

2.3.3 Protein Expression - Overnight cultures of E. coli BL21 (DE3)pLysS or Rosetta-

gami B (DE3)pLysS transformed with either pET19b-PSI or pET32z-PSI were used to

express PSI as per the manufacturer’s instructions. Cells were harvested by centrifugation

at 4,500 × g for 10 min at 4° C and stored at –20° C until further use. Frozen cells were

thawed at room temperature (RT) and resuspended in 20 mL of 20 mM Tris-Cl pH 7.5.

Suspensions were incubated at RT for 1 h with gentle shaking and the resulting cell lysates

were centrifuged at 21,000 × g for 30 min at 4o C to remove insoluble matter.

2.3.4 Protein Purification - The following applies only to Trx-PSI fusion protein

purification since pET19b-derived PSI was expressed at far lower concentrations and thus

was not pursued to purity. Protein purification was performed using an AKTATM FPLC

system (GE Healthcare, Piscataway, NJ, USA). Cell lysate soluble fractions were applied

to five 1 mL u-MAC columns in series (EMD Biosciences, San Diego, CA, USA)

equilibrated with 300 mM NaCl / 20 mM imidazole in 50 mM sodium phosphate pH 7.4

(binding buffer), followed by washing with the same buffer until a steady baseline was

obtained. Samples were eluted with 300 mM NaCl / 250 mM imidazole in 50 mM sodium

phosphate pH 7.4, then dialyzed in 20 mM Tris-Cl pH 7.4. Thrombin was added to the

30

dialysates at a 1:2000 mass ratio for incubation at RT for at least 12 h followed by re-

application of samples to u-MAC in binding buffer 3 times consecutively at 2 mL/min to

remove the Trx fusion tag. Flow-through was collected, dialyzed as above, then applied to

a 1 mL MonoQ column (GE Healthcare, Piscataway, NJ, USA) and separated using a 0-

500 mM NaCl gradient in 10 mM Tris-Cl pH 7.4. Eluent sample was further purified and

desalted on a 3 mL RPC column (GE Healthcare, Piscataway, NJ, USA), washed with 2%

acetonitrile/0.065% TFA and eluted with a 90 mL gradient (80% acetonitrile/0.05% TFA

elution buffer). The PSI peak, verified by SDS-PAGE and amino acid analysis (Advanced

Protein Analysis Center, The Hospital for Sick Children, Toronto, ON, Canada), was

collected and placed under vacuum in a Centrivap (Labconco Corp., Kansas City, MO,

USA) for 1 h at RT to remove the majority of the acetonitrile followed by dialysis against

4 x 1 L of 5 mM Tris-Cl pH 7.4 using 1 kDa MWCO dialysis tubing.

2.3.5 SDS-PAGE - Tris-glycine buffered SDS-PAGE was conducted according to the

method of Laemmli (Laemmli, 1970) in a Mini-Protean III electrophoresis cell (Bio-Rad,

Hercules, CA, USA). Gels were stained with GelCode Blue® (Pierce Biotechnology Inc.,

Rockford, IL, USA), and were analyzed for band size and relative intensities using a

ChemiGenius II system (Perkin Elmer, Waltham, MA, USA).

2.3.6 Crystallization - The purified StAP PSI protein sample was crystallized using

the sitting-drop vapor-diffusion method at 293 K using Qiagen PEG Suite screen solutions.

The best crystals appeared in the drop containing 0.4 µL protein solution / 0.2 µL reservoir

solution, equilibrated against 75 µL reservoir solution (0.2 M lithium sulfate/20%

PEG3350).

31

2.3.7 Diffraction Data Collection, Structure Solution and Refinement - X-ray

diffraction data for StAP PSI crystals were collected to 1.9 Å resolution using a Rigaku

MicroMax 007HF rotating anode and a MAR345dtb system at a wavelength of 1.5418 Å.

A data set was collected at 100 K using 25% (v/v) glycerol added to the reservoir solution

as cryo-protectant. All data sets were indexed and integrated using the program XDS

(Kabsch, 1993). Integrated intensities were converted to structure factors with modules

F2MTZ and CAD of CCP4 (Collaborative Computational Project, 1994). BUCCANEER

(Cowtan, 2006) was used for initial automated model building. The structure was refined

with REFMAC5 (Murshudov et al, 1997), rebuilt with COOT (Emsley & Cowtan, 2004),

and analyzed using PROCHECK (Laskowski et al, 1993) and COOT. Structural

superpositions were performed using SSM (Krissinel & Henrick, 2004) and ALIGN

(Cohen, 1997). Figures were generated using PYMOL (DeLano, 2002) and UCSF Chimera

(Pettersen et al, 2004).

2.3.8 Circular Dichroism (CD) Spectropolarimetry - CD analysis of PSI secondary

structure was carried out using a Jasco J-810 spectropolarimeter (Jasco Inc., Easton, MD,

USA). 200 µL of 200 µg/mL PSI was loaded into a 1 mm pathlength quartz cell and

scanned over 180-260 nm at 100 nm/min, 0.5 s response, standard sensitivity, and RT.

Buffers 140 mM NaCl / 10 mM Tris-Cl pH 7.4 or 140 mM NaCl / 20 mM MES pH 4.5

were degassed under vacuum. For reducing condition effect determinations, DTT was

added to final concentrations of 1.0, 2.5 and 5.0 mM, and heating was at 95º C in a standard

heating block for 5 min in a fume hood, followed by a 30 min cooling period on the bench

top.

32

2.3.9 Preparation of Large Unilamellar Vesicles (LUV) - LUVs were made of

equimolar phosphatidylcholine (PC), phosphatidylethanolamine (PE) and/or phosphatidyl-

serine (PS). To obtain 4 mM phospholipid (PL) suspensions, aliquots of 12.5 mg/mL PL

stocks were mixed in a tube and dried under N2 flush for at least 30 min, then suspended

in 500 µL of 80 mM calcein / 140 mM NaCl / 10 mM HEPES pH 7.4 by incubation at 37º

C with periodic sonication and vortexing over a minimum of 30 min. LUVs were prepared

using a standard mini-extruder (Avanti Polar Lipids Alabaster, AL, USA) containing a 100

nm pore membrane. The LUV prep was then desalted to remove untrapped calcein by gel

filtration using a 5 mL HiTrapTM desalting column (GE Healthcare, Piscataway, NJ, USA)

and visual detection of free calcein in column. To quantify PL post-desalting, the micro-

Bartlett phosphorous assay (Bartlett, 1959; Fiske & Subbarow, 1925) was used to

determine the concentration of PLs based on inorganic phosphate content (Sharom et al,

1995). Vesicle concentrations were calculated using PL concentrations, average vesicle

diameter (140 nm), and the previously reported areas per lipid molecule; 59.7 Å2 for PC

(Petrache et al, 1998), 57.4 Å2 for PS, and 59.2 Å2 for PE (Huster et al, 2000).

2.3.10 LUV Disruption Assays - PSI-caused perturbation of LUVs was measured by

calcein leakage (MacDonald et al, 1991; Matsuzaki et al, 1989) as detected using a Victor2

1420 Multilabel Counter (Perkin Elmer, Waltham, MA, USA) at 25° C. 200 µL reactions

were set up in 96-well microplates with varying concentrations of LUVs, 500 nM PSI and

either 140 mM NaCl / 10 mM HEPES pH 7.4 or 140 mM NaCl / 20 mM MES pH 4.5.

Leakage was detected using excitation at 385 nm and emission at 435 nm with 3 s shaking

between readings. End points were measured by incubating LUVs in 0.5% triton for each

33

condition. Non-linear regression analyses were done using GraphPad Prism 4 (GraphPad

Software Inc., La Jolla, CA, USA).

2.3.11 Atomic Force Microscopy - A Veeco Picoforce Multimode Scanning Probe

Microscope was used with a Nanoscope IV controller to image PE:PS membranes on the

native oxide layer on silicon substrates. Images were collected in contact mode before and

after the addition of PSI protein in situ. Suspensions of 3 mg/mL PE:PS vesicles were

incubated on substrates for 60 min, rinsed with 100 L 140 mM NaCl/20 mM MES pH 4.5

to remove unfused material, and inserted into the fluid cell under 50 L of buffer. Soft

triangular cantilevers were used with spring constants between 0.02 and 0.03 N/m, and the

force applied during each scan was 1.5-2.0 nN. Scans of 5 m 5 m were collected at a

rate of 1.5 Hz, and 10 m 10 m scans were collected at a rate of 0.75 Hz, corresponding

to a tip velocity of 15.2 m/s. After scanning the same region of substrate repeatedly over

30 min, 20 L of 25 M protein solution was injected directly into the buffer in the fluid

cell, resulting in 7 M PSI. Successive images were generated for a single region for time-

lapse data, and for unscanned regions at the end of incubation, to assess changes to the

membrane that were caused by repeated scanning of the AFM tip.

2.3.12 Particle Size Determination by Light Scattering - LUVs (100 μM) at pH 4.5

were incubated with PSI at RT and subjected to light scattering in a Malvern Zetasizer

Nano-S (Malvern Instruments, Malvern, Worcestershire, UK). A standard 1 mL cuvette

was used containing 0.6 mL sample that was allowed to equilibrate for a minimum of 15

min. Three consecutive measurements of five 30 s runs each were averaged using the

refractive index for polystyrene, yielding the calculated average sizes and polydispersity

indices.

34

2.4 Results

2.4.1 Structure Solution and Refinement - Recombinant StAP PSI was expressed and

purified to >98% purity with a typical yield of 5 mg/L culture and its identity was verified

by N-terminal sequencing (Advanced Protein Analysis Centre, Toronto, ON, Canada).

Diffraction data collection statistics are presented in Table 2.1. Crystals were hexagonal in

space group P3221 with unit cell parameters of a=b=56.47, c=55.34 Å. The Matthews

coefficient (Matthews, 1968) for the crystals was 2.24 Å3 Da-1, assuming the presence of

one molecule in the asymmetric unit. StAP PSI exhibits a high level of sequence identity

(53%) with the SAPLIP domain of prophytepsin PSI (PDB code 1QDM; residues 4S to

102S) which was used as a model for molecular replacement automated search by

PHASER (Read, 2001). The starting model consisted of a compact molecule and produced

a weak solution. Analysis of the initial map showed that PHASER had placed only half of

the initial model in a proper orientation despite good quality of the resulting electron

density; however, there was sufficient density to accommodate the properly oriented half.

At the next step, BUCCANEER (Cowtan, 2006) was used for automated model building,

thereby producing the model of the StAP PSI structure with proper side chains for residues

59-100 and assigning the other residues as polyalanine. Iterative refinement of the partial

model using REFMAC5 (Murshudov et al, 1997) and rebuilding in the electron density

maps using COOT (Emsley & Cowtan, 2004) produced the next model which

corresponded to an elongated, boomerang-shaped molecule. When all the residues visible

in the electron density were built, TLS parameters were introduced during the refinement.

The overall anisotropy was modeled with TLS parameters by dividing the molecule into

three groups comprising residues 0-26, 27-82, and 83-103. The final model lacks residues

35

Table 2.1: Data collection and refinement statistics

A. Data collection statisticsa

Space group P3221

Unit cell parameters a, b, c (Å) 56.47, 56.47, 55.34

Temperature (K) 100

Wavelength (Å) 1.5418

Resolution (Å) 40.0-190 (2.00-1.90)

Rmerge (%)b 5.9 (90.0)

Completeness (%) 99.9 (100.0)

I/ 24.10 (2.87)

Unique reflections 8355 (1159)

Redundancy 10.32 (10.27)

No. of molecules/asymmetric unit 1

B. Refinement statistics

Resolution (Å) 25.0-1.90

Working set: number of reflections 7936

Rfactor (%) 18.7

Test set: number of reflections 417

Rfree (%) 24.9

Protein atoms 633

No. of water molecules 63

C. Geometry statistics

rmsd (bond distance) (Å) 0.02

rmsd (bond angle) (deg.) 1.81

Ramachandran plotc

Most favored region (%) 98.6

Additionally allowed regions (%) 1.4

Generously allowed regions (%) 0.0

Disallowed regions (%) 0.0

D. PDB code 3RFI

a The values in parentheses are for the highest resolution shell b Rmerge = h i |<I>h – I h,i| / hi Ih,i. c As defined by PROCHECK.

36

40-63 which could not be built due to insufficient electron density in this region of the

crystal. The statistics of refinement for the refined structure are presented in Table 2.1, and

atomic coordinates and structure factors have been deposited in the PDB with the code

3RFI.

2.4.2 Tertiary and Quaternary Structures of StAP PSI - The overall fold of StAP PSI

has a boomerang shape with an extended, open conformation (Fig. 2.1.A) composed of

four helices labeled H1-H4; residues 1-24, 27-34, 66-82 and 85-99, respectively. H1 is

connected to H4 via a disulfide linkage formed between Cys6 and Cys99, H3 is cross-

linked to H2 via a disulfide bond between Cys31 and Cys71, and Cys37 forms a disulfide

bond with Cys68. The tertiary structure is organized in such a way that one side (top) of

the molecule is enriched with polar residues and the other side (bottom) is enriched with

hydrophobic residues (Fig. 2.1B-C). Due to the crystallographic symmetry, two molecules

form a very tight dimer (Fig. 2.1B-C) with a buried surface area of 1746 Å2, with the

residues involved in the formation of the dimer interface being predominantly

hydrophobic.

2.4.3 pH Dependence of Secondary Structure - Since PSI-induced membrane

disruption requires acidic conditions (Egas et al, 2000), the secondary structures of StAP

PSI at neutral and acidic pH were compared. CD scans were done in the same buffers used

for all other experiments, i.e. 140 mM NaCl buffered by either 10 mM Tris-Cl (pH 7.4) or

20 mM Na-MES (pH 4.5). Data below 195 nm was noisy so secondary structure content

could not be quantified. Qualitatively, the scans revealed spectra typical for high helix

proteins (Sreerama et al, 1999); distinct negative absorption peaks occurred in the 220 nm

37

Figure 2.1: The structure of StAP PSI. (A) Stereo view of the StAP PSI monomeric structure,

shown in ribbon representation. (B) Crystallographic dimer of StAP PSI; two molecules are shown

as ribbons inside the transparent surface of the dimer. (C) Electrostatic surface representation of

the StAP PSI structure.

38

and 208 nm spectral regions (Fig. 2.2). Helix content was higher overall at pH 4.5, similar

to the pH-dependence of influenza A hemagglutinin fusion peptide (Han et al, 2001).

2.4.4 Secondary Structure Dependence on Disulfide bonds - The structure of StAP PSI

contains three disulfide bonds within its relatively small 12 kDa tertiary structure and it

was recently suggested that these cross-links are critical to PSI antimicrobial function

(Muñoz et al, 2010). Hence, the dependence of PSI secondary structure on the presence of

cystines was investigated. Fig. 2.3 contains the spectra of PSI at three concentrations of the

reducing agent DTT, each with and without heating. No changes were observed for PSI

heated at 95º C under non-reducing conditions indicating that native PSI secondary

structural elements were heat stable. DTT resulted in high interference at wavelengths

below 200 nm so its levels were limited to 5 mM or lower. Qualitatively from Fig. 2.3, a

relatively minor loss of helix structure (220 nm and 208 nm negative peaks) occurred with

the presence of reducing agent in an apparent dose-dependent manner. By contrast, when

PSI was heated under reducing conditions, a more pronounced loss of secondary structure

was observed resulting in an approximate two-state structure change such that heating in

2.5 mM DTT resulted in a more dramatic loss of secondary structure. Further doubling of

reducing agent concentration did not cause an equivalent effect suggesting differential

susceptibility of the respective disulfide bonds; the more robust cystine(s) were apparently

critical to secondary structure stability of this predominantly helical structure.

2.4.5 Membrane Disruption Activity – Vesicle Leakage - Large unilamellar vesicles

(LUV) containing self-quenching fluorophore (80 mM calcein) were used as substrate for

the characterization of StAP PSI phospholipid bilayer disruption activity. Four

combinations of PLs were tested at varying concentrations. Neutral vesicles made of 1:1

39

Figure 2.2: Effect of acidification on StAP PSI secondary structure. Far-UV CD spectra

in 140 mM NaCl at pH 4.5 (red squares) and pH 7.4 (shaded circles).

40

Figure 2.3: Importance of disulfide bonds on StAP PSI secondary structure. Far-UV

CD spectra of StAP PSI at varying concentrations of reducing agent DTT with (dotted lines) and

without (solid lines) heating. The locations of the various disulfide bonds are indicated (red) in the

accompanying structure (bottom).

41

PC:PE were not affected by StAP PSI at any concentration tested and none of the PL

combinations were disrupted at neutral pH. By contrast, equimolar preparations of PS

combined with PC and/or PE resulted in readily detectable activity and leakage rates were

calculated for 0.5 μM PSI over the PL concentration range 20-500 μM. Accurate LUV

disruption rate determinations above 500 μM were prevented by excessive fluorescence

signal and the quality of determinations below 20 μM were limited by excessive

background noise levels due to untrapped calcein. Rate determinations were calculated

relative to PL concentrations in units of μM/min (in terms of both PL and LUV

concentration) and yielded initial rates with low relative errors. PSI-induced lipid bilayer

disruption varied non-linearly with PL / LUV concentration and the order of leakage rates

was PE:PS > PC:PS > PE:PC:PS within the concentration range tested (Fig. 2.4).

Non-linear regression analyses for both one-phase exponential association and the

Michaelis-Menten equation were compared by F-tests using GraphPad Prism 4 and the

independent data sets for all three PL combinations fit better to the Michaelis-Menten

model. Table 2.2 summarizes the kinetic results including goodness of fit parameters for

the regression analyses. Since the mechanism of action has not been characterized (see the

Discussion section), apparent Michaelis constants (Km) were not reported since their

meaning in terms of substrate affinity would be undefined in the absence of understanding

the relative dissociation rates of PSI and PL in original bilayer (reactant; k-1) and displaced

PL in new environment (product; k2). A lack of data at higher lipid concentrations (which

produced excessive fluorescence signals) resulted in a relatively large turnover number

standard error for PC:PS. LUV concentrations above 500 μM would allow for more precise

42

Figure 2.4: Kinetics of LUV disruption by 0.5 µM StAP PSI at 25º C. comparison of

three acidic phospholipid mixtures (activity against non-acidic PE:PC was not detectable). Two

vertical axes are presented for disruption rates in terms of both phospholipid concentration (left

axis) and vesicle concentration (right axis).

43

Table 2.2: Turnover and goodness of fit to the Michaelis-Menten model for StAP PSI-

induced vesicle leakage. Data are shown relative to both phospholipid and vesicle

concentrations.

Phospholipid kcat apparent (min-1) kcat apparent (×10-4 min-1) R2

Composition (phospholipid) (vesicle)

PC:PS 71 ± 30 3.6 ± 2 0.992

PE:PS 45 ± 3 2.3 ± 0.2 0.999

PE:PC:PS 25 ± 4 1.3 ± 0.2 0.996

44

kinetic parameter determinations and will require further characterization. Nevertheless,

the overall fit to the model was good for each data set including that for PC:PS (R2=0.99).

2.4.6 Membrane Disruption Activity – Atomic Force Microscopy - AFM height images

of PE:PS bilayers were collected using contact mode as described above in the

Experimental Procedures. LUVs were incubated on the surface of the native oxide layer on

a silicon wafer in the same buffer used for LUV disruption assays for 1 h resulting in bilayer

fusion to the substrate surface. The fused membrane was organized as 5 nm high islands

that were approximately 100 nm wide, dispersed across the substrate (Fig. 2.5.A; top). The

bilayer height was within the expected range (Alattia et al, 2006; Egawa & Furusawa, 1999)

and repeated scans in contact mode over a single region did not change the bilayer

morphology significantly. Upon PSI injection, membrane patches fused to form larger

regions of uniform membrane separated by larger membrane-free areas (Fig. 2.5.A;

bottom). Additionally, 20-50 nm high lipid islands formed in regions unperturbed by the

AFM tip (Fig. 2.5.B; right), thus PSI appeared to induce fusion of membrane patches as

evidenced by bilayer rearrangement in areas that were not repeatedly scanned. Such

formations did not occur on regions subjected to successive scanning; on these regions,

considerable smoothing of the membrane occurred. PSI apparently softened or lubricated

bilayers allowing them to be displaced by the AFM tip upon repeated scanning. Fig. 2.5.B

(right) shows the distinct effects of PSI both with (centre square region) and without

(surrounding region) repeated AFM tip scanning. Large regions of 5 nm high bilayer were

observed after 8 scans on a single region whereas repeated scanning in the absence of PSI

did not result in similar bilayer rearrangement.

45

Figure 2.5: AFM height images of PE:PS bilayer patches at pH 4.5 on the native oxide

layer of a silicon wafer. (A) Top: Successive scans without PSI (left to right); the membrane is

patchy with ~100 nm wide islands of height 5 nm. Bottom: Successive scans with PSI (left to right);

the membrane is smooth, transforming from patchy islands to large continuous membrane. The

white lines indicate height sections shown below each image. (B) AFM height images of the same

region as (A); left: pre-injection, middle: 50 min (8th scan) post-injection, right: 60 min (9th scan)

post-injection showing a larger region zoomed out to twice the scan width. Note the smoothing of

the membrane over the region repeatedly scanned by the AFM tip post-injection and the appearance

of islands much taller (white) than the original 5 nm bilayer height in the region not affected by

repeated scanning of the AFM tip.

46

2.4.7 Membrane Disruption Activity – Light Scattering - LUVs (100 μM) at pH 4.5

were incubated with PSI at RT and particle size was measured by light scattering in a

Malvern Zetasizer Nano-S (Malvern Instruments, Malvern, Worcestershire, UK); results

are summarized in Table 2.3. After 60 min at a PSI concentration equivalent to LUV

disruption assays (0.5 μM), average particle size slightly increased and a concomitant

increase in the polydispersity index (mass distribution) was more pronounced (Table 2.3).

By 100 min, average particle size had increased by nearly two thirds (65%), and the

polydispersity index had more than tripled. Higher PSI concentration (4 μM) was assayed

to further investigate the peak area tendency toward higher vesicle size. After just 15 min

equilibration time post-PSI addition, average particle size and polydispersity index had not

only increased more than for the entire 0.5 μM PSI assay, but a newly formed, larger vesicle

size (1510 nm) accounting for one third of the total peak area was observed. Together with

the above AFM results, vesicle size measurements indicated that PSI activity resulted in

lipid bilayer fusion.

47

Table 2.3: Effect of StAP PSI on LUV size at pH 4.5.

Time

(min)

Z-average size

(nm) Polydispersity Index

0.5 µM

0 142 ± 0.85 0.0870 ± 0.020

60 146 ± 1.6 0.172 ± 0.018

100 235 ± 0.20 0.300 ± 0.0030

4 µM

0 137 ± 1.2 0.0737 ± 0.0098

15

(overall) 257 ± 4.4 0.529 ± 0.064

15 (peak 1) 167 ± 11 * 65% ± 1 of total area

15 (peak 2) 1510 ±300 * 34% ± 1 of total area

48

2.5 Discussion

2.5.1 Structural Comparison - The overall fold of StAP PSI is similar to the open,

extended form of human saposin C (Fig. 2.6.A) as observed in the tetragonal crystal

structure (Rossmann et al, 2008). The structure of StAP PSI has been compared with the

PSI domain of plant phytepsin (Kervinen et al, 1999) revealing that the RMSD for 38

aligned Cα atoms is 0.9 Å. Overall, superposition of a pseudo-monomer StAP PSI structure

(constructed from its dimer crystal structure) onto the PSI of phytepsin produced an RMSD

of 1.4 Å (71 Cα pairs). Helices H1 and H4 of StAP PSI were superimposable with the

equivalent helices of phytepsin PSI (Fig. 2.6.A) while helices H2 and H3 from the second

StAP PSI molecule in the crystallographic dimer superimpose onto helices H2 and H3 of

phytepsin PSI. This result clearly shows that the crystallographic dimer structure of StAP

PSI is composed of two extended, domain-swapped monomers. The StAP PSI structure

has also been compared with three different crystal structures of human saposin C (Fig.

2.6.A). The Cα atoms of residues 6-19 from saposin C were superimposed (LSQ) to the Cα

atoms of the first 13 residues of the StAP PSI structure yielding an RMSD of 0.6 Å

(tetragonal crystal form, 2Z9A), 0.6 Å (orthorhombic crystal form, 2QYP) and 0.9 Å

(hexagonal crystal form, 2GTG). This “open” conformation for SAPLIPs has also been

observed for saposin B (1N69), and all three structures are shown side-by-side in Fig.

2.6.B.

2.5.2 Saposin-like Activity - StAP PSI-liposome disruption rates varied with PL/LUV

concentration in a non-linear fashion and, to our knowledge, the present study is the first

to characterize a SAPLIP’s kinetic activity for varying PL/LUV concentrations of multiple

PL compositions, revealing Michaelis-Menten-like kinetics. In attempting to elucidate the

49

Figure 2.6: Tertiary structural features of StAP PSI. (A) Structural superposition of StAP

PSI (green) with the plant-specific insert domain of prophytepsin (cyan), tetragonal (yellow),

hexagonal (grey) and orthorhombic (red) crystal forms of saposin C. (B) Side by side comparison

of the “open” saposin boomerang fold of saposin B (1N69), saposin C (2QYP) and StAP PSI

(3RFI). (C) Pairwise structural alignment of influenza A hemagglutinin fusion peptide (1IBN;

purple) superimposed over StAP PSI (3RFI; green) using UCSF Chimera version 1.5.2

implementing the Needleman-Wunsch algorithm. The overall RMSD was 2.013 Å for 19 residues

that aligned, ignoring gaps.

50

nature of PSI activity, AFM results suggested that StAP PSI caused pronounced

rearrangement of acidic PL bilayers under the same conditions used for liposome

disruption. The type of lipid structures that formed were similar to those previously

characterized for saposin C which showed that saposin C interactions with PL bilayers

(You et al, 2003) resulted in rearrangement of membrane patches of increased thickness

and membrane destabilization (Alattia et al, 2006; You et al, 2003). Saposin C has also

been shown to cause vesicle fusion (Vaccaro et al, 1994) which involves insertion of the

terminal helices into the membrane where saposin–membrane and saposin–saposin

interactions carry out a ‘clip-on’ mechanism at neutral and acidic pH. Saposin C causes

fusion when present at as little as 0.05 µM and produced fused vesicles up to 3000 nm

(Wang et al, 2003).

By contrast, saposin B displays fusogenic activity only against anionic vesicles and

exclusively at acidic pH, and induces only minor vesicle size increases at protein

concentrations of 1 µM or below. However, at 2 µM saposin B induces more significant

vesicle fusion yielding an average product ~1800 nm in diameter (Wang et al, 2003). The

fusogenic results for StAP PSI in the present study were strikingly similar in that 0.5 µM

PSI induced only minor average diameter increases whereas 4 µM PSI induced dramatic

changes resulting in new lipid structures averaging 1510 nm (Table 2.3). Also similar to

StAP PSI disruption and AFM observations, saposins B and D disrupt anionic membranes

in a pH-dependent process where they solubilize (Ciaffoni et al, 2001; Remmel et al, 2007)

and mobilize (Remmel et al, 2007) lipids. Furthermore, saposin B binds and transfers PLs

of anionic membranes such that it has a preference for PC transfer (Ciaffoni et al, 2006).

Although StAP PSI-mediated vesicle disruption rates were all higher for PE:PS within the

51

concentration range used in the present study, kinetic analysis suggested a higher maximum

velocity for PC:PS disruption (P=0.06), indicating a PL preference similar to saposin B.

Perhaps the bulkier choline substituent, relative to ethanolamine, results in more favorable

PSI interactions with the bilayer surface and/or results in different bilayer packing density.

A study of the nature of SAPLIP PL preference is underway.

The crystal structure of saposin B with bound lipid indicated that PLs interact with the

dimeric form of the protein. Dimerization occurs via clasping together two V-shaped

protein monomers, thereby forming a shell-like monolayer of α-helices with a long

interface that buries a relatively large hydrophobic cavity (Ahn et al, 2003). A similar V-

shape and quaternary structural arrangement was observed in the present study for the StAP

PSI crystal structure, although PSI and open saposin C contain a more obtuse angle in their

boomerang shape relative to saposin B (see Fig. 2.6.B), according to structural alignments.

Furthermore, PSI membrane effects as detected by AFM were in partial agreement with

saposin B in that new lipid structures formed that were higher than the surrounding

bilayers; structures observed for saposin B were smaller in size however (Alattia et al,

2006). These newly formed ‘granules’ could be dislodged by the scanning AFM tip

indicating that they were loosely bound upon saposin B action (Alattia et al, 2006).

Similarly, raised lipid bilayers acted upon by StAP PSI became more fluid in regions

repeatedly scanned by the AFM tip and were spread out into smooth continuous lipid

regions (Fig. 2.5.B) indicating solubilising and/or mobilizing activity.

2.5.3 Plant AP PSI-Activity - Recombinant StAP PSI is toxic to plant (P. infestans and

F. solani) and human (S. aureus, B. cereus and E. coli) pathogens, and it permeabilizes

their plasma membranes (Muñoz et al, 2010). In both procardosin A and StAP, the PSI has

52

been confirmed to be responsible for membrane interactions (Egas et al, 2000; Muñoz et

al, 2010). The interaction of procardosin A PSI with membrane vesicles is pH-dependent

and varies with lipid composition (Egas et al, 2000), in agreement with the present study.

Interestingly, the only other crystal structure for a plant AP PSI, prophytepsin, was shown

to have a tertiary structure consisting of the “closed” saposin fold (Kervinen et al, 1999).

That crystal structure dealt with a PSI attached to its parent zymogen molecule and

contrasts with the “open” saposin fold observed in the present study for a PSI independent

of its zymogen source. By definition, fusogenic activity via the “clip-on” mechanism

(Wang et al, 2003) is dependent upon open structure dimerization (interfacing of the

respective hydrophobic surfaces). Thus, we propose that plant AP PSIs likely change from

closed to open fold upon release from the parent AP molecule thereby facilitating bilayer

interaction and subsequent protein quaternary structure dimerization, yielding fusion of

neighboring bilayer structures.

The cytotoxicity and plasma membrane interactions of StAP PSI were previously shown

to be dependent on its secondary and/or tertiary structure, as evidenced by loss of activity

upon DTT treatment (Muñoz et al, 2010). In the present study, CD scans at 1 mM DTT

resulted in no apparent spectral change, and only minor changes were indicated at DTT

concentrations up to 5 mM (without heating). This suggested that disulfide bonds are not

critical to PSI secondary structure under normal temperature conditions. When synthetic

peptides equivalent to the individual helices of saposin C (i.e., no native tertiary structure)

were studied, bilayer fusogenic activity was not observed (Wang et al, 2003), suggesting a

critical role of tertiary structure in SAPLIP-catalyzed bilayer fusion. Since cystines are

53

critical to antimicrobial function (Muñoz et al, 2010), their role must be to maintain tertiary

structure required for fusogenic activity.

Additionally, the DTT titration CD experiment indicated that the disulfide bonds

conferred stability to PSI secondary structures as evidenced by a much more pronounced

CD spectral change for heat-treated PSI. The latter was superimposable with that for the

non-reducing, unheated sample (Fig. 2.3) indicating that the PSI disulfide bonds apparently

protected the individual helices from heat denaturation. Since both open and closed forms

of SAPLIPs contain the same disulfide bonds (Bruhn, 2005), the primary role of PSI

cystines is not likely the maintenance of the overall fold, but perhaps to confer rigidity.

Fusion of adjacent membranes consists of the displacement of lipid from its stable,

energetically favorable bilayer environment to an aqueous, high energy intermediate state

and such a transition requires enzymatic action (Baker & Agard, 1994). Perhaps the energy

required to catalyze this event is related to multimer formation (Markovic et al, 1998) and

conformation changes (Chernomordik et al, 1998; Kingsley et al, 1999) related to PSI

rigidity/stability dependent on disulfide bonds.

2.5.4 Fusogenic Mechanism - There is a correlation between the ability of a fusion

peptide to adopt a helical configuration and its ability to promote membrane fusion

(Hernandez et al, 1996). In addition to saposins, the formation of dimers via complexing

of stable, predominantly helical structures is a protein structure scheme common to yeast

SNARE-mediated membrane fusion (Paumet et al, 2005; Paumet et al, 2004) as well as

viral fusases which are derived from larger precursors that require proteolytic processing

to potentiate their fusion activity (Hernandez et al, 1996). These associations of α-helices

contain one hydrophobic face, an arrangement similar to StAP PSI with its 5 helices,

54

hydrophobic inner cavity and N-terminus. An important fusase fusion peptide for virus-

cell fusion is the N-terminal portion of influenza A hemagglutinin which adopts an α-

helical conformation in lipid bilayers (Han & Tamm, 2000), constitutes an autonomous

folding unit in the membrane and catalyzes lipid exchange between juxtaposed membranes

(Han et al, 2001).

In some plant APs, the release of PSI occurs via proteolytic cleavage upon acid-induced

autoactivation of the precursor protein and subsequent processing, albeit via self-cleavage

(Faro et al, 1999; Glathe et al, 1998; Ramalho-Santos et al, 1998). Interestingly, influenza

A hemagglutinin fusion peptide is inaccessible to membranes at neutral pH; however, a

drop of the pH inside the endosome below a critical threshold induces a large

conformational change in the parent protein and is subsequently activated by a protease

(plasmin) that cleaves the precursor polypeptide (Lazarowitz et al, 1973) into two disulfide-

linked polypeptides and a fusion peptide (Gray et al, 1996). The hemagglutinin fusion

peptide has a slightly higher helix content at pH 5 than at pH 7.4 as revealed by comparison

of CD spectra (Han et al, 2001; Han & Tamm, 2000) which are remarkably similar to those

for StAP PSI in the present study (Fig. 2.2) in terms of overall shape (dominant helix

content), as well as their x-intercepts and superimposed relative spectra (gain of helical

structure upon acidification). In addition, the fold of the hemagglutinin fusion peptide (Han

et al, 2001) is similar to the StAP PSI structure reported here as well as to saposins (Vaccaro

et al, 1995): Structural alignment of the N-terminus of hemagglutinin fusion peptide

(1IBN) (Han et al, 2001) with the N-terminus of StAP PSI revealed similar (RMSD 1.43

Å for 9 Cα carbons) helix-kink-helix folds (Fig. 2.6.C). The tryptophan within the

hemagglutinin fusion peptide has been shown to induce its characteristic boomerang shape

55

(Lai et al, 2006) which is critical to the fusogenic and membrane disrupting activities of

this fold (Lai & Tamm, 2010). Alignment of StAP PSI with hemagglutinin fusion peptide

suggested a similar role for the critical tryptophan residue (Fig. 2.6.C). In this context, the

likely protein structural reason for the acidic pH requirement of AP PSI-membrane

interactions is the existence of acid-induced helical structure critical for membrane

interaction.

2.5.5 Fusase Within a Protease - The apparent sharing of a common hydrophobic

region fold that is subject to similar pH-dependent secondary structure changes required

for membrane fusion suggests a common mode of action. Perhaps this helix-kink-helix fold

is universal (i.e., animal; saposins, virus; hemagglutinin, plant; APs) in its membrane

fusogenic nature extending functionality across various species and kingdoms.

Collectively, the findings that a plant AP domain displays fusase-like activity (liposome

disruption, bilayer solubilization/lubrication and bilayer fusion) as well as fusase-like

structure-function character (inter- and intramolecular helix oligomer association,

hemagglutinin fusion peptide-like fold, hydrophobic helix end region, and pH-dependence

of secondary structure-activity) lead to the conclusion that they are indeed fusase-like

proteins, acting as discrete entities.

Considering the myriad proteins that have more than one function, the idea of one gene

- one protein - one function is insufficient in the study of proteins (Jeffery, 2009). Recently,

a bifunctional AP has been engineered (Bryksa et al, 2010), however, no reports have

characterized cases of AP “moonlighting” (Moore, 2004), i.e., to serve an additional

function beyond the main enzymatic reaction (Copley, 2003), and only one moonlighting

plant peptidase (mitochondrial processing peptidase) has been reported (Braun & Schmitz,

56

1999; Moore, 2004). In this case, the multiple functions arise from a singular structural

fold (Braun & Schmitz, 1995) whereas PSIs are structurally unrelated to their AP “hosts”.

That is, the PSI domain of a plant AP has apparent enzymatic activity and is independent

of its “parent” proenzyme which has its own distinct class of enzymatic activity. Thus, the

saposin-like domains of plant APs present a unique case: A distinct, functionally unrelated

domain within the primary structure of another domain (the C-terminal domain) of its

enclosing protein. Fusase activity from within a protease sequence presents, to our

knowledge, the first confirmation and characterization of an independently acting “enzyme

within an enzyme”.

57

2.6 References

Ahn, V. E., Faull, K. F., Whitelegge, J. P., Fluharty, A. L. & Prive, G. G. (2003) Crystal

structure of saposin B reveals a dimeric shell for lipid binding. Proceedings of the

National Academy of Sciences of the United States of America, 100(1), 38-43.

Ahn, V. E., Leyko, P., Alattia, J.-R., Chen, L. & Privé, G. G. (2006) Crystal structures of

saposins A and C. Protein Science, 15(8), 1849-1857.

Alattia, J. R., Shaw, J. E., Yip, C. M. & Prive, G. G. (2006) Direct visualization of saposin

remodelling of lipid bilayers. Journal of Molecular Biology, 362(5), 943-953.

Anderson, D. H., Sawaya, M. R., Cascio, D., Ernst, W., Modlin, R., Krensky, A. &

Eisenberg, D. (2003) Granulysin crystal structure and a structure-derived lytic

mechanism. Journal of Molecular Biology, 325(2), 355-365.

Baker, D. & Agard, D. A. (1994) Influenza hemagglutinin: Kinetic control of protein

function. Cell, 2(10), 907-910.

Bartlett, G. R. (1959) Phosphorous assay in column chromatography. Journal of Biological

Chemistry, 234(3), 466-468.

Bissonnette, M. L. Z., Donald, J. E., DeGrado, W. F., Jardetzky, T. S. & Lamb, R. A.

(2009) Functional analysis of the transmembrane domain in paramyxovirus F protein-

mediated membrane fusion. Journal of Molecular Biology, 386(1), 14-36.

Blundell, T. & Johnson, M. (1993) Catching a common fold. Protein Science, 2(6), 877-

883.

Braun, H.-P. & Schmitz, U. K. (1995) Are the ‘core’ proteins of the mitochondrial bc1

complex evolutionary relics of a processing protease? Trends in Biochemical Sciences,

20(5), 171-175.

Braun, H.-P. & Schmitz, U. K. (1999) The protein-import apparatus of plant mitochondria.

Planta, 209(3), 267-274.

Bruhn, H. (2005) A short guided tour through functional and structural features of saposin-

like proteins. Biochemical Journal, 389, 249-257.

Bryksa, B. C., Bhaumik, P., Magracheva, E., DeMoura, D. C., Kurylowicz, M., Zdanov,

A., Dutcher, J. R., Wlodawer, A. & Yada, R. Y. (2011) Structure and mechanism of the

58

saposin-like domain of a plant aspartic protease. Journal of Biological Chemistry,

286(32), 28265-28275.

Bryksa, B. C., Horimoto, Y. & Yada, R. Y. (2010) Rational redesign of porcine pepsinogen

containing an antimicrobial peptide. Protein Engineering Design & Selection, 23(9),

711-719.

Chernomordik, L. V., Frolov, V. A., Leikina, E., Bronk, P. & Zimmerberg, J. (1998) The

pathway of membrane fusion catalyzed by influenza hemagglutinin: Restriction of

lipids, hemifusion, and lipidic fusion pore formation. Journal of Cell Biology, 140(6),

1369-1382.

Ciaffoni, F., Salvioli, R., Tatti, M., Arancia, G., Crateri, P. & Vaccaro, A. M. (2001)

Saposin D solubilizes anionic phospholipid-containing membranes. Journal of

Biological Chemistry, 276(34), 31583-31589.

Ciaffoni, F., Tatti, M., Boe, A., Salvioli, R., Fluharty, A., Sonnino, S. & Vaccaro, A. M.

(2006) Saposin B binds and transfers phospholipids. Journal of Lipid Research, 47(5),

1045-1053.

Cohen, G. E. (1997) ALIGN: A program to superimpose protein coordinates, accounting

for insertions and deletions. Journal of Applied Crystallography, 30(6), 1160-1161.

Collaborative Computational Project, N. (1994) The CCP4 suite: Programs for protein

crystallography. Acta Crystallographica, D50, 760-763.

Copley, S. D. (2003) Enzymes with extra talents: Moonlighting functions and catalytic

promiscuity. Current Opinion in Chemical Biology, 7(2), 265-272.

Cowtan, K. (2006) The Buccaneer software for automated model building. 1. Tracing

protein chains. Acta Crystallography, D62(9), 1002-1011.

Davies, D. (1990) The structure and function of the aspartic proteinases. Annual Reviews

in Biophysics and Biophysical Chemistry, 19, 189-215.

DeLano, W. L. (2002) The PyMOL Molecular Graphics System. DeLano Scientific.

Egas, C., Lavoura, N., Resende, R., Brito, R. M., Pires, E., de Lima, M. C. & Faro, C.

(2000) The saposin-like domain of the plant aspartic proteinase precursor is a potent

inducer of vesicle leakage. Journal of Biological Chemistry, 275(49), 38190-38196.

Egawa, H. & Furusawa, K. (1999) Liposome adhesion on mica surface studied by atomic

force microscopy. Langmuir, 15(5), 1660-1666.

59

Emsley, P. & Cowtan, K. (2004) Coot: model-building tools for molecular graphics. Acta

Crystallographica, D60, 2126-2132.

Faro, C., Ramalho-Santos, M., Vieira, M., Mendes, A., Simoes, I., Andrade, R., Verissimo,

P., Lin, X.-l., Tang, J. & Pires, E. (1999) Cloning and characterization of cDNA

encoding cardosin A, an RGD-containing plant aspartic proteinase. Journal of

Biological Chemistry, 274(40), 28724-28729.

Fiske, C. H. & Subbarow, Y. (1925) The colorimetric determination of phosphorous.

Journal of Biological Chemistry, 66(2), 375-400.

Frazão, C., Bento, I., Costa, J., Soares, C. M., Veríssimo, P., Faro, C., Pires, E., Cooper, J.

& Carrondo, M. A. (1999) Crystal structure of cardosin A, a glycosylated and Arg-Gly-

Asp-containing aspartic proteinase from the flowers of Cynara cardunculus L. Journal

of Biological Chemistry, 274(39), 27694-27701.

Ghannoum, M. A. & Rice, L. B. (1999) Antifungal agents: Mode of action, mechanisms

of resistance, and correlation of these mechanisms with bacterial resistance. Clinical

Microbiology Reviews, 12(4), 501-517.

Glathe, S., Kervinen, J., Nimtz, M., Li, G. H., Tobin, G. J., Copeland, T. D., Ashford, D.

A., Wlodawer, A. & Costa, J. (1998) Transport and activation of the vacuolar aspartic

proteinase phytepsin in barley (Hordeum vulgare L.). Journal of Biological Chemistry,

273(47), 31230-31236.

Gordon, L. M., Lee, K. Y. C., Lipp, M. M., Zasadzinski, J. A., Walther, F. J., Sherman, M.

A. & Waring, A. J. (2000) Conformational mapping of the N‐terminal segment of

surfactant protein B in lipid using 13C‐enhanced Fourier transform infrared

spectroscopy. Journal of Peptide Research, 55(4), 330-347.

Gray, C., Tatulian, S. A., Wharton, S. A. & Tamm, L. K. (1996) Effect of the N-terminal

glycine on the secondary structure, orientation, and interaction of the influenza

hemagglutinin fusion peptide with lipid bilayers. Biophysical Journal, 70(5), 2275-

2286.

Guevara, M. G., Oliva, C. R., Huarte, M. & Daleo, G. R. (2002) An aspartic protease with

antimicrobial activity is induced after infection and wounding in intercellular fluids of

potato tubers. European Journal of Plant Pathology, 108(2), 131-137.

60

Han, X., Bushweller, J. H., Cafiso, D. S. & Tamm, L. K. (2001) Membrane structure and

fusion-triggering conformational change of the fusion domain from influenza

hemagglutinin. Nature Structural Biology, 8, 715-720.

Han, X. & Tamm, L. K. (2000) A host–guest system to study structure–function

relationships of membrane fusion peptides. Proceedings of the National Academy of

Sciences of the United States of America, 97(24), 13097-13102.

Hanahan, D. (1983) Studies on transformation of Escherichia coli with plasmids. Journal

of Molecular Biology, 166(4), 557-580.

Hernandez, L. D., Hoffman, L. R., Wolfsberg, T. G. & White, J. M. (1996) Virus-cell and

cell-cell fusion. Annual Review of Cell and Developmental Biology, 12, 627-661.

Huster, D., Arnold, K. & Gawrisch, K. (2000) Strength of Ca2+ binding to retinal lipid

membranes: Consequences for lipid organization. Biophysical Journal, 78(6), 3011-

3018.

Jeffery, C. J. (2009) Moonlighting proteins - an update. Molecular BioSystems, 5(4), 345-

350.

Kabsch, W. (1993) Automatic processing of rotation diffraction data from crystals of

initially unknown symmetry and cell constants. Journal of Applied Crystallography, 26,

795-800.

Kasson, P. M., Lindahl, E. & Pande, V. S. (2010) Atomic-resolution simulations predict a

transition state for vesicle fusion defined by contact of a few lipid tails. PLoS

Computational Biology, 6(6), e1000829.

Kervinen, J., Tobin, G. J., Costa, J., Waugh, D. S., Wlodawer, A. & Zdanov, A. (1999)

Crystal structure of plant aspartic proteinase prophytepsin: inactivation and vacuolar

targeting. EMBO Journal, 18(14), 3947-3955.

Kingsley, D. H., Behbahani, A., Rashtian, A., Blissard, G. W. & Zimmerberg, J. (1999) A

discrete stage of baculovirus GP64-mediated membrane fusion. Molecular Biology of

the Cell, 10(12), 4191-4200.

Kolter, T. & Sandhoff, K. (2005) Principles of lysosomal membrane digestion: Stimulation

of sphingolipid degradation by sphingolipid activator proteins and anionic lysosomal

lipids. Annual Review of Cell and Developmental Biology, 21(1), 81-103.

61

Krissinel, E. & Henrick, K. (2004) Secondary-structure matching (SSM), a new tool for

fast protein structure alignment in three dimensions. Acta Crystallographica Section D:

Biological Crystallography, 60(12), 2256-2268.

Laemmli, U. K. (1970) Cleavage of structural proteins during the assembly of the head of

bacteriophage T4. Nature, 227, 680-685.

Lai, A. L., Park, H., White, J. M. & Tamm, L. K. (2006) Fusion peptide of influenza

hemagglutinin requires a fixed angle boomerang structure for activity. Journal of

Biological Chemistry, 281(9), 5760-5770.

Lai, A. L. & Tamm, L. K. (2010) Shallow boomerang-shaped influenza hemagglutinin

G13A mutant structure promotes leaky membrane fusion. Journal of Biological

Chemistry, 285(48), 37467-37475.

Laskowski, R. A., MacArthur, M. W., Moss, D. S. & Thornton, J. M. (1993) PROCHECK:

A program to check the stereochemical quality of protein structures. Journal of Applied

Crystallography, 26, 283-291.

Lazarowitz, S. G., Goldberg, A. R. & Choppin, P. W. (1973) Proteolytic cleavage by

plasmin of the HA polypeptide of influenza virus: Host cell activation of serum

plasminogen. Virology, 56(1), 172-180.

Liepinsh, E., Andersson, M., Ruysschaert, J. M. & Otting, G. (1997) Saposin fold revealed

by the NMR structure of NK-lysin. Nature Structural Biology, 4(10), 793-795.

Lukatsky, D. B. & Frenkel, D. (2004) Multiple stalk formation as a pathway of defect-

induced membrane fusion. The European Physical Journal E: Soft Matter and

Biological Physics, 14(1), 3-6.

MacDonald, R. C., MacDonald, R. I., Menco, B. P., Takeshita, K., Subbarao, N. K. & Hu,

L.-r. (1991) Small-volume extrusion apparatus for preparation of large, unilamellar

vesicles. Biochimica et Biophysica Acta - Biomembranes, 1061(2), 297-303.

Markovic, I., Pulyaeva, H., Sokoloff, A. & Chernomordik, L. V. (1998) Membrane fusion

mediated by baculovirus gp64 involves assembly of stable gp64 trimers into

multiprotein aggregates. Journal of Cell Biology, 143(5), 1155-1166.

Matsuda, J., Vanier, M. T., Saito, Y., Tohyama, J., Suzuki, K. & Suzuki, K. (2001) A

mutation in the saposin A domain of the sphingolipid activator protein (prosaposin) gene

62

results in a late-onset, chronic form of globoid cell leukodystrophy in the mouse. Human

Molecular Genetics, 10(11), 1191-1199.

Matsuzaki, K., Harada, M., Handa, T., Funakoshi, S., Fujii, N., Yajima, H. & Miyajima,

K. (1989) Magainin 1-induced leakage of entrapped calcein out of negatively-charged

lipid vesicles. Biochimica et Biophysica Acta - Biomembranes, 981(1), 130-134.

Matthews, B. W. (1968) Solvent content of protein crystals. Journal of Molecular Biology,

33(2), 491-491-497.

Mazorra-Manzano, M. A., Tanaka, T., Dee, D. R. & Yada, R. Y. (2010) Structure–function

characterization of the recombinant aspartic proteinase A1 from Arabidopsis thaliana.

Phytochemistry, 71(5-6), 515-523.

Meyer, V. (2008) A small protein that fights fungi: AFP as a new promising antifungal

agent of biotechnological value. Applied Microbiology and Biotechnology, 78(1), 17-

28.

Moore, B. (2004) Bifunctional and moonlighting enzymes: Lighting the way to regulatory

control. Trends in Plant Science, 9(5), 221-228.

Murshudov, G. N., Vagin, A. A. & Dodson, E. J. (1997) Refinement of macromolecular

structures by the maximum-likelihood method. Acta Crystallographica, 53(3), 240-255.

Mutlu, A. & Gal, S. (1999) Plant aspartic proteinases: Enzymes on the way to a function.

Physiologia Plantarum, 105(3), 569-576.

Muñoz, F. F., Mendieta, J. R., Pagano, M. R., Paggi, R. A., Daleo, G. R. & Guevara, M.

G. (2010) The swaposin-like domain of potato aspartic protease (StAsp-PSI) exerts

antimicrobial activity on plant and human pathogens. Peptides, 31(5), 777-785.

Paumet, F., Rahimian, V., Di Liberto, M. & Rothman, J. E. (2005) Concerted auto-

regulation in yeast endosomal t-SNAREs. Journal of Biological Chemistry, 280(22),

21137-21143.

Paumet, F., Rahimian, V. & Rothman, J. E. (2004) The specificity of SNARE-dependent

fusion is encoded in the SNARE motif. Proceedings of the National Academy of

Sciences of the United States of America, 101(10), 3376-3380.

Payie, K. G., Tanaka, T., Gal, S. & Yada, R. Y. (2003) Construction, expression and

characterization of a chimaeric mammalian-plant aspartic proteinase. Biochemical

Journal, 372(3), 671-678.

63

Payie, K. G., Weadge, J. T., Tanaka, T. & Yada, R. Y. (2000) Purification, N-terminal

sequencing and partial characterization of a novel aspartic proteinase from the leaves of

Medicago sativa L. (alfalfa). Biotechnology Letters, 22(19), 1515-1520.

Petrache, H. I., Tristram-Nagle, S. & Nagle, J. F. (1998) Fluid phase structure of EPC and

DMPC bilayers. Chemistry and physics of lipids, 95(1), 83-94.

Pettersen, E. F., Goddard, T. D., Huang, C. C., Couch, G. S., Greenblatt, D. M., Meng, E.

C. & Ferrin, T. E. (2004) UCSF Chimera -A visualization system for exploratory

research and analysis. Journal of Computational Chemistry, 25(13), 1605-1612.

Qiao, H., Armstrong, R. T., Melikyan, G. B., Cohen, F. S. & White, J. M. (1999) A specific

point mutant at position 1 of the influenza hemagglutinin fusion peptide displays a

hemifusion phenotype. Molecular Biology of the Cell, 10(8), 2759-2769.

Ramalho-Santos, M., Veríssimo, P., Cortes, L., Samyn, B., Van Beeumen, J., Pires, E. &

Faro, C. (1998) Identification and proteolytic processing of procardosin A. European

Journal of Biochemistry, 255(1), 133-138.

Read, R. J. (2001) Pushing the boundaries of molecular replacement with maximum

likelihood. Acta Crystallographica Section D: Biological Crystallography, 57(10),

1373-1382.

Remmel, N., Locatelli-Hoops, S., Breiden, B., Schwarzmann, G. & Sandhoff, K. (2007)

Saposin B mobilizes lipids from cholesterol-poor and bis(monoacylglycero)phosphate-

rich membranes at acidic pH. FEBS Journal, 274(13), 3405-3420.

Rossmann, M., Schultz-Heienbrok, R., Behlke, J., Remmel, N., Alings, C., Sandhoff, K.,

Saenger, W. & Maier, T. (2008) Crystal structures of human saposins C and D:

Implications for lipid recognition and membrane interactions. Structure, 16(5), 809-817.

Runeberg‐Roos, P., Tormakangas, K. & Östman, A. (1991) Primary structure of a barley‐

grain aspartic proteinase. European Journal of Biochemistry, 202(3), 1021-1027.

Schaller, A. & Ryan, C. A. (1996) Molecular cloning of a tomato leaf cDNA encoding an

aspartic protease, a systemic wound response protein. Plant Molecular Biology, 31(5),

1073-1077.

Sharom, F. J., DiDiodato, G., Yu, X. & Ashbourne, K. J. D. (1995) Interaction of the P-

glycoprotein Multidrug Transporter with Peptides and Ionophores. Journal of

Biological Chemistry, 270(17), 10334-10341.

64

Simoes, I. & Faro, C. (2004) Structure and function of plant aspartic proteinases. FEBS

Journal, 271(11), 2067-2075.

Sreerama, N., Venyaminov, S. Y. U. & Woody, R. W. (1999) Estimation of the number of

α‐helical and β‐strand segments in proteins using circular dichroism spectroscopy.

Protein Science, 8(2), 370-380.

Stiasny, K., Kössl, C., Lepault, J., Rey, F. A. & Heinz, F. X. (2007) Characterization of a

Structural Intermediate of Flavivirus Membrane Fusion. PLoS Pathogens, 3(2), e20.

Tseng, H.-K. & Perfect, J. R. (2011) Strategies to manage antifungal drug resistance.

Expert Opinion on Pharmacotherapy, 12(2), 241-241-256.

Vaccaro, A. M., Salvioli, R., Barca, A., Tatti, M., Ciaffoni, F., Maras, B., Siciliano, R.,

Zappacosta, F., Amoresano, A. & Pucci, P. (1995) Structural analysis of saposin C and

B. Journal of Biological Chemistry, 270(17), 9953-9960.

Vaccaro, A. M., Tatti, M., Ciaffoni, F., Salvioli, R., Serafino, A. & Barca, A. (1994)

Saposin C induces pH-dependent destabilization and fusion of phosphatidylserine-

containing vesicles. FEBS Letters, 349(2), 181-186.

Wang, Y., Grabowski, G. A. & Qi, X. (2003) Phospholipid vesicle fusion induced by

saposin C. Archives of Biochemistry and Biophysics, 415(1), 43-53.

You, H. X., Qi, X., Grabowski, G. A. & Yu, L. (2003) Phospholipid Membrane Interactions

of Saposin C: In Situ Atomic Force Microscopic Study. Biophysical Journal, 84(3),

2043-2057.

Zhai, Y. & Saier, M. H. (2000) The amoebapore superfamily. Biochimica et Biophysica

Acta - Reviews on Biomembranes, 1469(2), 87-99.

65

Chapter 3: Protein Structure Insights into the Bilayer

Interactions of the Saposin-Like Domain of Solanum

tuberosum Aspartic Protease

Note: The content of this chapter was submitted (December 2016) to The Journal of

Biological Science for review.

3.1 Abstract

The present study sought to clarify the nature of plant saposin membrane-active features

through a detailed examination of the Solanum tuberosum aspartic protease saposin-like

domain structure. Cryo-transmission electron microscopy revealed phospholipid bilayer

vesicle morphologies having prominent flat surfaces. Furthermore, the plant saposin was

active exclusively as an acid pH-dependent dimer. Secondary structure changes occurred

between inactive and active pH conditions, and consisted of a 7% gain in helix and

concomitant 7% loss in strand/turn. Upon mixing with bilayer vesicles, increases in helix

structure and tryptophan fluorescence emission occurred on a similar timescale. Reduction

of disulfide bonds did not appear to alter secondary structure at low pH, however, it did

result in faster apparent vesicle fusion. Individual PSI segments corresponding to its

secondary structural elements (four helices and an unstructured mid-sequence region) were

subsequently characterized. Although inactive, helix-1 (the N-terminal 24 residues) was

the only segment which experienced pH-dependent secondary structure changes, becoming

disordered at neutral pH. Helix-3 (equivalent to saposin helix-1) was the only peptide

segment that showed activity, causing both vesicle leakage and apparent fusion. Helix-3

secondary structure was pH-insensitive and predominantly β-strand/turn. Furthermore,

mutation of the sole positively charged residue (Lys83Gln) on helix-3, disordered its

66

secondary structure and eliminated bilayer interaction. Taken together, the helical character

of the component segments of the plant saposin domain are not intrinsically determined

and are instead dependent upon overall tertiary and quaternary structures. Implications of

the findings for the bilayer disruption mechanism(s) are discussed for both the saposin

domains as well as helix-3.

3.2 Introduction

Unlike their non-plant counterparts, many plant aspartic proteases (APs) contain a ~100-

residue sequence insert that results in an ‘extra’ domain attached to the C-terminal lobe.

This unique additional segment, aptly referred to as the plant-specific insert (PSI), is a

member of the saposin-like protein (SAPLIP) superfamily (Bruhn, 2005; Bryksa et al,

2011; Kervinen et al, 1999; Mutlu & Gal, 1999; Runeberg‐Roos et al, 1991). Delineation

of the PSI primary sequence revealed its N- and C-terminal sequence portions to occur in

the opposite order, or swapped, relative to other SAPLIPs, hence the term swaposin

(Runeberg‐Roos et al, 1991). Upon post-translational processing and activation, plant AP

saposin domains are either removed (heterodimeric plant APs) or retained (monomeric

plant APs) (Egas et al, 2000; Glathe et al, 1998; Kervinen et al, 1999; Ramalho-Santos et

al, 1998). Even in the latter scenario, the PSI and AP moieties exist as structurally distinct

features of a shared overall tertiary structure (Frazão et al, 1999; Kervinen et al, 1999), and

function independently (Bryksa et al, 2011; Frazão et al, 1999; Muñoz et al, 2010; Pereira

et al, 2013; Tormakangas et al, 2001).

SAPLIPs have diverse biological functions (Anderson et al, 2003; Bruhn, 2005;

Liepinsh et al, 1997; Matsuda et al, 2001; Zhai & Saier, 2000) all involving lipid

67

interactions (Bruhn, 2005; Egas et al, 2000; Kolter & Sandhoff, 2005). As such, plant AP

saposin-like domains display bilayer perturbation (Bryksa et al, 2011; Egas et al, 2000;

Muñoz et al, 2011) and membrane permeabilizing (Mendieta et al, 2006; Muñoz et al,

2014) activities. Mechanisms by which membrane-SAPLIP contacts occur have been

proposed spanning a variety of modes (Alattia et al, 2007; Andreu et al, 1999; Miteva et

al, 1999; Qi & Grabowski, 2001), with PSI being poorly understood relative to the more

studied members of the SAPLIP superfamily such as helminth amoebapores, mammalian

saposins A-D, surfactant protein B, NK-lysin and granulysin. The roles of PSIs in vivo have

been proven to not only include vacuolar targeting and Golgi/endosomal sorting (Glathe et

al, 1998; Kervinen et al, 1999; Mutlu & Gal, 1999; Runeberg‐Roos et al, 1991), but also a

clear association with the ability of some plants to stave off phytopathogen invasion

(Guevara et al, 2002). Globally, the negative impacts on agricultural products by plant

diseases cause annual losses in the tens of billions (USD) (Almgren et al, 2000; Edwards

et al, 1993; Montesinos & Bardají, 2008). Due to their roles in plant pathogen resistance

(Guevara et al, 2002), and senescence and post-harvest physiology (Payie et al, 2000;

Schaller & Ryan, 1996), plant APs are important for the development of novel biocontrol

technologies (Curto et al, 2014) as well as understanding biochemical and biological

phenomena of natural plant disease resistance.

Biochemical insights into the antimicrobial etiology of Solanum tuberosum aspartic

protease (StAP) PSI have been revealed (Mendieta et al, 2006; Muñoz et al, 2011; Pagano

et al, 2007) including its effectiveness against human pathogens and inhibition of

phytopathogen sporulation via interaction with, and permeabilization of, microbial plasma

membranes (Muñoz et al, 2010). The PSI domain of cirsin, from Cirsium vulgare (Lufrano

68

et al, 2012), has also been shown to have activity against different phytopathogens in vitro

(Curto et al, 2014). Furthermore, cirsin has been produced in a biologically active form

using a generally regarded as safe yeast recombinant expression system (Curto et al, 2014),

a promising advance towards a potential role in pest and/or pathogen control. In order to

develop novel plant biocontrol agents as well as therapeutic drugs for the treatment of

fungal infections (Ghannoum & Rice, 1999) in immunocompromised patients (Meyer,

2008; Tseng & Perfect, 2011), characterizing the structural features that modulate lipid

bilayer interactions is critical for rational design approaches.

Our group previously reported on StAP PSI structure and function including the findings

that the domain-swapped plant saposin adopted the less commonly observed “open”

saposin fold consisting of a dimer of extended V-shaped monomers (Bryksa et al, 2011).

It was also revealed by direct contact imaging that PSI rendered anionic phospholipid

bilayer more prone to deformation, and that PSI secondary structure is pH-dependent, but

independent of its disulfide bonds. The PSI was also shown to cause apparent fusion of

anionic phospholipid (PL) bilayer liposomes (Bryksa et al, 2011). Our findings led us to

conclude that, due to multiple structure-function commonalities with certain fusases, PSI

may possess a fusion peptide-type mechanism of action that is common to influenza A

hemagglutinin fusion peptide. Furthermore, we predicted the existence of an acid pH-

induced helical structure within PSI as being the structural basis for the low pH requirement

of PSI-membrane interactions (Bryksa et al, 2011).

The present investigation sought to better understand the structural underpinnings for

PSI-mediated bilayer interactions and structure changes. Specifically, what significance, if

any, do the tertiary (open V-shape) and quaternary (homodimer) structures observed in the

69

PSI crystal structure (3RFI) have for the pH-dependence of PSI-bilayer interactions? This

knowledge is required for ultimately assessing the applicability of the clip-on mechanism

originally postulated for saposin C (Wang et al, 2003) as well as aiding in identifying what

region of the PSI structure mediates bilayer contact and selectivity for anionic targets? To

answer these questions, secondary, tertiary and quaternary structures were characterized

for intact PSI as well as its individual component regions (based upon 3RFI; see Figure

3.1) in terms of bilayer disruption, and vesicle fusion and morphological changes.

3.3 Materials and Methods

3.3.1 Materials - A PSI synthetic gene optimized for expression in E. coli was purchased

from Mr. Gene GmbH (Regensburg, Germany). Plasmid pET32b(+), E. coli Rosetta-gami

B (DE3)pLysS, and u-MACTM columns were obtained from EMD Biosciences (San Diego,

CA, USA). E. coli TOP10F’ was from Invitrogen (San Diego, CA, USA). GenEluteTM

Plasmid Miniprep Kit was obtained from Sigma-Aldrich (St. Louis, MO, USA). The

QIAquick® PCR Purification Kit and QIAquick® Gel Extraction Kit were from Qiagen

(Germantown, MD, USA). Restriction enzymes, T4 DNA ligase and Pfu DNA polymerase

were obtained from Fermentas (Burlington, ON, Canada). Primers were synthesized by

Sigma Genosys (Oakville, ON, Canada) and thrombin was purchased from Fisher

Scientific (Ottawa, ON, Canada). The RPC column was from GE Healthcare (Piscataway,

NJ, USA). Phospholipids were from Avanti Polar Lipids (Alabaster, AL, USA).

3.3.2 PSI Expression and Purification – PSI was recombinantly expressed and purified

as per (Bryksa et al, 2011).

70

Figure 3.1: Crystal structure of StAP PSI (3RFI). The figure indicates the positions of the

three disulfide bonds in red, the two Trp residues in blue, the sequences of the five chosen segments

(H1, H2, X, H3, H4), and the three ionizable residues of H3 (shown in purple). Note that “H” stands

for helix, however “X” is an unknown structure that was not resolvable in the 3RFI crystal structure.

71

3.3.3 Preparation of Large Unilamellar Vesicles - LUV stocks were prepared as per

(Bryksa et al, 2011) with the exception that 80 mM calcein / 140 mM NaCl / 25 mM Na-

acetate pH 4.5 was used in order to suspend dried PL mixtures at 42º C after removal of

storage solvent under nitrogen flush. Phospholipids used were 1-palmitoyl-2-oleoyl-sn-

glycero-3-phosphocholine (POPC), 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoethanol-

amine (POPE) and 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoserine (POPS). Note that

information on vesicle stability is summarized in Appendix 3.A.

3.3.4 Circular Dichroism Spectropolarimetry - CD analysis of PSI secondary structure

was carried out using a Jasco J-810 spectropolarimeter (Jasco Inc., Easton, MD, USA).

Samples (200 µL of 100 µg/mL PSI or 200 µg/mL peptide) were scanned over 180-260

nm at 100 nm/min, 0.5 s response, standard sensitivity, and at ambient temperature using

a 1 mm pathlength quartz cell. Iso-ionic (164 mM) Na-phosphate or Na-acetate saline

buffers were degassed under vacuum. Secondary structure contents were calculated using

DICHROWEB (Whitmore & Wallace, 2008) with SELCON3 (Sreerama et al, 1999) and

CDSSTR (Sreerama & Woody, 2000; 2004) algorithms. For CD time trial scans of PSI

mixed with LUVs, 4 µM (47 µg/mL) or 10 µM PSI 118 µg/mL) and 100 nm LUVs (100

µM total PL; 1:1:1 molar ratio of POPC:POPE:POPS) were scanned over 196 nm -260 nm

at 200 nm/min every 2 min.

3.3.5 LUV Disruption Assays - PSI-caused perturbation of LUVs was measured by

calcein leakage (56, 57) as detected using a Victor2 1420 Multilabel Counter (Perkin

Elmer, Waltham, MA, USA) at 25° C. 200 µL reactions were set up in 96-well microplates

with varying concentrations of LUVs, 0.5-2.0 µM StAP PSI, or 4-10 µM peptides (peptide

concentrations up to 40 µM were screened to verify lack of activity for H1, H2, H1H2 and

72

H4) in either Na-phosphate or Na-acetate buffered saline (constant ionic strength 164 mM).

Leakage was detected using excitation at 385 nm and emission at 435 nm with 3 s shaking

between readings. Leakage was monitored until increase in fluorescence ceased and/or for

a minimum of 15 min / maximum 120 min. End points were measured by incubating LUVs

in 0.5% triton for each condition. Non-linear regression analyses were done using

GraphPad Prism 4 (GraphPad Software Inc., La Jolla, CA, USA).

3.3.6 Bilayer Fusion Assays - Ten μM PSI or peptide was incubated with 100 nm LUVs

(100 μM total PL; 1:1 molar ratio of POPE:POPS) at 22° C in either 25 mM Na-acetate /

140 mM NaCl pH 4.5 or 5 mM Na-phosphate / 140 mM NaCl pH 7.4 (control). Mixtures

were then monitored for changes in average LUV size by dynamic light scattering in a

Malvern Zetasizer Nano-S (Malvern Instruments, Malvern, Worcestershire, UK) using a

disposable polystyrene 1.5 mL semi-microcuvette. Three consecutive measurements of

five runs (30 s per run) were averaged using the refractive index for polystyrene.

Considerations for using DLS to characterize PSI/peptide-induced vesicle effects are

discussed in Appendix B.

3.3.7 Sedimentation Equilibrium Analytical Centrifugation - Sedimentation

equilibrium studies were carried out using a Beckman Optima XL-A Analytical

Ultracentrifuge (Biomolecular Interactions and Conformations Facility, University of

Western Ontario, Canada). An An-60Ti rotor and six-channel cells with Eponcharcoal were

used at 22,000 rpm, 28,000 rpm and 35,000 rpm, 20° C. To achieve adequate detectability,

26-28 µM PSI were compared at iso-ionic strength in 27 mM sodium phosphate pH 3.0,

20 mM sodium phosphate pH 6.2 or 10 mM sodium phosphate pH 7.4, each containing

140 mM NaCl. Absorbance measurements at 280 nm were collected in 0.002 cm radial

73

steps and averaged over 10 readings. Solvent densities (ρ) were calculated using

SEDNTERP software. Partial specific volume (ν) of the protein was calculated from its

amino acid sequence to be 0.13337 mL/g. Data were analyzed using a single ideal species

model in GraphPad Prism.

3.3.8 Tryptophan Intrinsic Fluorescence Emission Spectrometry - Fluorescence

emission spectra were recorded using a Shimadzu RF-540 spectrofluorophotometer

(Shimadzu Corporation, Kyoto, Japan) with a 1-cm quartz three-sided ultra-micro cuvette

in a water-circulating, temperature-controlled cell holder at 25° C.  The settings used were

λexcitation 295 nm with 3 nm slit width and λemission scan 300−400 nm with 3 nm slit width.

PSI samples were diluted to a working stock concentration of 80 µM and allowed to

incubate in a water bath for 15 min at 25° C as was done for a working stock of anionic

LUVs (1000 µM total PLs) to pre-equilibrate. Ten µM PSI and 100 µM PL (1:1:1 molar

ratio of POPC:POPE:POPS as 100 nm LUVs) were mixed in buffered saline pH 4.5 and

monitoring began at 20 s. Scans were completed at 500 nm/min allowing for a scan every

15 s.

3.3.9 Cryo-Transmission Electron Microscopy - Liposome imaging was done at The

Microscopy Imaging Facility, University of Guelph Advanced Analysis Centre (Guelph,

Canada) on an FEI Tecnai G2 F20 TEM (FEI Co., Hillsboro, Oregon, USA) with a bottom

mount Gatan 4k CCD camera and 200kV field emission. Samples consisted of 1000 µM

or 500 µM PL (as 100 nm LUV; 1:1 molar ratio of POPE:POPS) with or without 10 µM

or 16 µM PSI in 25 mM Na-Acetate pH 4.5 / 140 mM NaCl incubated for 15 min at 22° C.

All were loaded onto sample grids and plunge flash frozen in a dust-free moisture-

controlled work space, and subsequently maintained at or below -170° C.

74

3.3.10 PSI Structure Component Peptides - Synthetic peptides corresponding to the

PSI structure regions (outlined in Figure 3.1) were purchased from GenicBio Ltd.

(Shanghai, China). The sequences were as follows (positively charged (bold underlined)

and aromatic (underlined italics) residues are indicated for easy reference):

H1: IVSMECKTIVSQYGEMIWDLLVSG

H2: VRPDQVCSQAGLCFV

H1H2: IVSMECKTIVSQYGEMIWDLLVSGVRPDQVCSQAGLCFV

X: DGAQHVSSNIKTVVERETEGSSVG

H3: EAPLCTACEMAVVWMQNQLKQ

H4: EGTKEKVLEYVNQLCEKIP

Peptides were delivered lyophilized in known quantities and were initially resuspended in

20 mM Na-phosphate pH 7.2 / 10 mM DTT (with volumes targeting 1 mg/mL peptide

concentration), and gently rocked at room temperature for 15 min followed by

centrifugation at 10,000 x g for 5 min. Only peptide H4 was stably soluble at this point. All

peptides were then further diluted 5-fold with 20 mM Na-phosphate pH 7.2 / 10 mM

followed by 2 h gentle rocking and clarification. Due to low solubility, H1 was prepared

alternatively by suspending directly into 6 mM NaOH followed by neutralization with 1 M

Na-phosphate pH 7.2 to a final concentration of 50 mM Na-phosphate. DTT was then

added to 10 mM final concentration. H1, H1H2, H3 and H4 were quantified by A280 and

stored at -30° C after flash freezing in liquid nitrogen as 50 µM stocks (except H1 which

was 20 µM). H2 and X could not be quantified accurately due to their negligible UV

absorptivities. Since neither peptide was used for quantitative purposes, and neither

presented solubility problems, their concentrations were based upon the weighed

lyophilized mass prior to solubilisation. Due to their respective solubility and quantitation

75

problems, H1H2 was also studied. Lastly, three H3 mutants (Glu63Gln, Glu74Gln and

Lys83Gln) were also purchased and prepared as above.

3.3.11 Statistical Analyses - Statistical significance of differences within and between

data sets were calculated with Kruskal-Wallis 1-way ANOVA and Dunn’s Multiple

Comparison Test using GraphPad Prism 6.07.

3.4 Results

3.4.1 StAP PSI Quaternary Structure in Solution - Although StAP PSI (3RFI)

crystalized as a dimer at neutral pH (Bryksa et al, 2011), its tertiary/quaternary structure

status in solution at or near its active pH remained unknown. With a predicted pI 4.6,

solubility of PSI close to optimal pH for bilayer disruption (Egas et al, 2000) is insufficient

for analytical centrifugation at the required concentration range. Our experience working

with PSI at pH 4.5 has been to work at or below 80 µM avoiding elevated centrifugal force

whereas it is soluble at ˃34 mg/mL at neutral pH. As a compromise, PSI was analyzed

below and above pH 4.5, at pH 3.0 and 6.2 as well as inactive conditions at pH 7.4.

Recombinant PSI was determined to have apparent masses of 12.5 and 13.6 kDa at pH 7.4

and 6.2, respectively, and approximately double, 21.7 kDa, at pH 3.0 (Figure 3.2). These

results indicated that recombinant PSI (11,771 Da) exists exclusively as a dimer in the

active pH range, and as a monomer near neutral, inactive pH. To determine the PSI

monomer-dimer status at pH 4.5, intrinsic tryptophan fluorescence emission was used

(Figure 3.3.A) which indicated it to be distinct from pH 6.2 and 7.4 (monomer), but similar

to pH 3.0 (dimer). Although the pH 3.0 dimer had stronger fluorescence emission, optimal

emission wavelength remained unchanged (Figure 3.3.B).

76

Figure 3.2: Sedimentation equilibrium analytical centrifugation analyses in iso-ionic

saline buffers at pH 3.0, 6.2 and 7.4. (A) Absorbance measurements at 280 nm collected in

0.002 cm radial steps and averaged over 10 readings; measurements were determined in triplicate.

(B) Summary of the calculated masses from matching tests (mean +/- one standard deviation, n=6).

77

Figure 3.3: Intrinsic Trp Fluorescence Emission. (A) Solution Intrinsic Trp Fluorescence

of StAP PSI at pH 3.0, 4.5, 6.2 and 7.4. (B) λmax at the indicated pH values were not significantly

different comparing all respective pairings (P>0.05). (C) Relative maximum fluorescence emission

were not significantly different (P>0.05) between pH 3.0 and 4.5, nor between pH 6.2 and 7.4

(P>0.05), whereas the former were each significantly different (P≤0.05) from the latter,

respectively, suggesting distinct electronic environments for one or both Trp residues at active and

inactive pH conditions. (D) Relative areas under the emission spectra were significantly different

(P≤0.05) for all data set pairings (mean +/- one standard deviation, n=3).

78

3.4.2 Secondary Structure Dependence on pH and Disulfide Bonds - PSI secondary

structure content was compared as a function of pH by far-UV circular dichroism (CD) in

iso-ionic buffers for non-saline, saline and reducing saline conditions (Figures 3.4.A-C,

respectively). From the non-saline spectra, a trend of increasing helix content with

decreasing pH (P≤0.05 for each paired comparison) and an accompanying increase in

strand content with increasing pH such that pH 3.0 < 4.5 < 6.2 (P≤0.05 for each paired

comparison) while pH 6.2 and 7.4 PSI were not significantly different (P>0.05). Unordered

secondary structure did not appear to follow clear patterns based on pH (Figure 3.4.D).

Spectra collected for reducing and non-reducing saline conditions were suitable only for

qualitative comparisons due to CD signal interference by DTT and NaCl at wavelengths

below 205 nm and 196 nm, respectively, as well as increasing UV absorbance by DTT

below 225 nm. The presence of saline conditions did not change the pattern of higher

helical character at lower pH values and increasing β-structure character with increasing

pH. However, a difference was noted in the saline spectra in that PSI appears to exist in

two discreet secondary structure states as evidenced by indistinguishable spectra pairings

at pH 3.0 and 4.5, and pH 6.2 and 7.4, respectively (Figures 3.4.A-B). This contrasted with

an overall pattern of continuous decreasing CD signal magnitudes over the whole spectra

for PSI in the absence of NaCl.

In addition to pH-dependence of secondary structure, differences in saline low pH

spectra relative to their non-saline counterparts (Figures 3.4.A-B) suggest that secondary

structure sensitivity across the pH range is mitigated by saline conditions. This is evidenced

by two observations: First, the magnitudes of the overall spectra are greater at pH 3.0 and

79

Figure 3.4: Far-UV CD spectra of StAP PSI. The spectra at pH 3.0 (blue), 4.5 (red), 6.2

(green) and 7.4 (purple) are for 100 µg/mL PSI and are averaged from five scans at RT. To maintain

consistent ionic strength between conditions, samples were prepared in either 12 mM sodium

phosphate pH 3.0, 25 mM sodium acetate pH 4.5, 10 mM sodium phosphate pH 6.2 or 5 mM

sodium phosphate pH 7.4 buffers. Spectra were collected for PSI in (A) buffer; (B) buffer

containing 140 mM NaCl; (C) Buffer containing 140 mM NaCl and 10 mM DTT, overlaid with

spectra from (B); and (D) Calculated secondary structure contents of StAP PSI at the indicated pH

values calculated using DICHROWEB (Whitmore and Wallace, 2008) and the SELCON3

(Sreerama et al., 1999) and CDSSTR (Sreerama and Woody, 2000) algorithms (calculated values

were averaged).

80

4.5; and second, there appears to be a distinct shift in the relative ellipticities for key

important local minima and maxima (see Figure 3.4.A,B,E). Dominant features particularly

evident in the pH 3.0 and 4.5 spectra include a large positive peak centred below 195 nm

and two narrow large negative peaks at 222 and 208 nm, indicators for high helix content

(Kelly & Price, 2000). Increasing pH resulted in reduced intensities of these peaks and a

broadening of the local minima in the 222-208 nm region (Figure 3.4.A), as reflected in

the calculated secondary structure contents (Figure 3.4.D) where helix structure is lost at

the expense of β-structure. The NaCl-containing samples’ spectra reveal qualitative

differences in that they have an overall broader and lower magnitude 222-208 nm region

at pH 3.0 and 4.5 (Figure 3.4.B). Although noise is limiting below 215 nm, the CD spectra

for reducing conditions suggest similar overall secondary structure (Figure 3.4.C).

3.4.3 pH-Dependence of PSI-induced Phospholipid Bilayer Disruption - Bilayer

disruption was compared at pH 3.0, 4.5, 6.2 and 7.4 using 1:1 POPE:POPS large

unilamellar vesicles (LUVs; 100 µM total PL) mixed with 0.5-10 µM PSI. Leakage was

monitored until increase in fluorescence ceased (15-120 min). Within roughly 15 min, PSI-

induced leakage consistently plateaued at approximately 25% of maximum fluorescence

(i.e., upon addition of 0.1% w/v Triton X-100). Results are summarized in Figure 3.5

showing that PSI caused vesicle leakage at pH 3.0 and 4.5, with the latter having a 3-fold

higher maximum leakage rate (significantly different, P≤0.05), while leakage was not

detected at pH 6.2 and 7.4 (not significantly different, P>0.05).

81

Figure 3.5: pH-dependence of PSI-induced LUV leakage. Maximum leakage rates in iso-

ionic saline buffers at 25° C are expressed as a function of fluorescence increase relative to fully

released fluorophore controls, i.e., Triton X-100-lysed LUVs.

82

3.4.4 Intrinsic Tryptophan Fluorescence for PSI in the Presence of Anionic Bilayer -

The intrinsic tryptophan fluorescence of PSI 1:1:1 POPC:POPE:POPS LUVs was

measured over a wide range of emission wavelengths (305–380 nm) and monitored over

time. Figure 3.6.A shows the emission scans monitoring PSI fluorescence for emission

wavelengths between 305-380 nm over 285 s. Note that the indicated times refer to scan

start times, each requiring 12 s to complete. Maximum fluorescence emission followed a

single phase non-linear association (R2=0.99), increasing steadily for approximately 1 min

followed by a deceleration period that was near static after 5 min (see Figure 3.6B).

However, λmax did not trend up or down (no significant departure from zero slope; P>0.05),

an unexpected finding as it was hypothesized that the N-terminal helix Trp would embed

in bilayer resulting in a blue shift in the emission spectrum.

3.4.5 PSI Secondary Structure in the Presence of Anionic Bilayer LUVs - PSI

secondary structure upon encountering anionic bilayer was monitored by collecting far-UV

CD spectra for 10 µM PSI with or without 1:1:1 POPC:POPE:POPS LUVs (100 µM PL)

in buffered saline pH 4.5 at 22° C. Continuous monitoring resulted in excessive UV

exposure such that coagulation of the LUV mixture into a gel-like state occurred within

minutes, therefore, spectra were collected at a relatively rapid scan rate. . Scans were

normalized, respectively, to control spectra for buffered-saline with or without LUVs

scanned in the same manner as samples in terms of UV exposure and timing in attempts to

account for any hidden UV-induced changes to PLs. PL-containing controls and samples

were also verified by dynamic light scattering (DLS) which confirmed that the respective

expected vesicle sizes were present after the elapsed time of the assays. Figure 3.7

summarizes the CD spectral changes with overlapping spectra beyond 8.5 min omitted for

83

Figure 3.6: Intrinsic Trp fluorescence emission of StAP PSI. (A) Spectra with anionic

bilayer vesicles - Times indicated in the legend correspond to scan start times (seconds) after

initial mixing of 10 µM PSI and 1:1:1 POPC:POPE:POPS LUVs (100 µM PL). (B) Maximum

fluorescence at λmax and λmax are summarized, indicating that no shift in λmax for Trp emission

occurs despite the increased emission over the period observed (mean +/- one standard deviation,

n=3).

84

Figure 3.7: Comparison of timed far-UV CD spectra of PSI with 1:1:1

POPC:POPE:POPS LUVs (100 µM PL) in buffered saline pH 4.5 at 22° C. Spectra

were collected through 30 min, however, overlapping scans are omitted for clarity. Each data set is

an average of 5 scans at 500 nm/min. The comparison is made to test the hypothesis that the PSI

would gain helix structure upon interacting with anionic bilayers.

85

clarity. The principal difference revealed by the scans was an increase in magnitude across

the 222-208 nm region at both PSI concentrations. At 222 and 208 nm, ellipticity increased

by approximately 26% and 16% for 10 µM PSI, and 13% and 4% for 4 µM PSI,

respectively, relative to the control signals at the respective wavelengths. Signal-noise was

only modestly increased by the presence of vesicles and increased absorbance of incident

light was only marginally higher below 215 nm, gradually increasing to ~130% of buffered

saline by 200 nm (data not shown).

3.4.6 Characterization of PSI-induced Bilayer Effects: Cryo-Transmission Electron

Microscopy - Imaging was used to characterize liposome morphology. Shown in Figure

3.8 are images representative of the observed phenomena acquired for two concentrations

of PSI+PL (10+500 µM and 16+1000 µM) having similar PSI:PL molar ratios (1:50 and

2:125, respectively) using 1:1 POPE:POPS LUVs, incubated at 22° C for 15 min prior to

cryo-plunging. Control LUVs were in agreement with LUV quality control DLS

measurements for stock LUV preparations, consistently yielding stable vesicles that

measured 125-140 nm. Note that vesicles substantially contacting the sample support grid

were excluded from analyses due to typical artefactual vesicle morphologies (Goodwin &

Khant). The images were collected from three separate experiments using different PSI

preparations and freshly made (<24 h) LUV preparations, verified for size by DLS. Drastic

changes were noted for PSI-treated LUVs which were categorized into five distinct

morphologies: (i) Less narrow oblong structures that appear to be collapsed vesicles (blue

arrows); (ii) Individual semi-circular liposomes with a flattened or straight edge (pink

arrows); (iii) Vesicles interfacing at respective flat edges (orange arrows);(iv) Wedge-type

vesicles having two flat edges meeting at their ends (yellow arrows); and (v) Narrow

86

Figure 3.8: Cryo-Transmission Electron Microscopy images of 1:1 POPE:POPS

LUVs in buffered saline at pH 4.5. Incubations were in 25 mM Na-Acetate / 140 mM NaCl

pH 4.5 for (A) untreated LUVs; (B) LUVs (500 µM PL) treated with 10 µM PSI; and (C) LUVs

(1000 µM PL) treated with 16 µM PSI. Assays were 15 min in length at 22° C at which time all

were cryo-plunged. Images are representative of the observed phenomena from 9 control and 8 test

images. Different observed morphologies are identified by arrows as follows: Oblong structures

that appear to be partially-collapsed vesicles (blue arrows); individual liposomes with a flattened

or straight edge (pink arrows); vesicles interfacing at respective flat edges (orange arrows); and

wedge-type vesicles having two flat edges meeting at an end (yellow arrows); and narrow elongated

rod-like structures approximately 100–200 nm in length (red arrows).

87

elongated rod-like structures approximately 100–200 nm in length (red arrows), which

appear to show evidence of bilayer by way of a thin, white, low contrast layer completely

surrounding a high density thin centre region, suggesting a collapsed liposome.

The various vesicle morphologies were enumerated and analyzed from all suitable (9

control and 8 test) cryo-TEM images (Figure 3.9). The occurrence and diameter of each

observed vesicle are summarized in Figure 3.9.A. Histograms of the vesicle diameter

distributions for circular vesicles exclusively as well as for overall vesicle populations are

summarized in Figure 3.9.B-C, respectively. Analyses of variance indicated that the

respective vesicle diameter distributions are significantly different (P≤0.05) for both

comparisons, respectively (Figures 3.9.B-C). Comparing the circular vesicle size

distributions (Figure 3.9.B), both 10 µM PSI/500 µM PL and 16 µM PSI/1000 µM PL tests

are significantly different (P<0.05) from control, respectively, as well as to each other (2-

tailed Kolmogorov–Smirnov tests). Comparing the vesicle diameter distributions of the

overall populations (i.e. all morphologies; Figure 3.9.C), the 10 µM PSI/500 µM PL

distribution is not significantly different (P˃0.05) from control while the 16 µM PSI/1000

µM PL distribution is significantly different (P≤0.05) from control.

The differences in diameter distributions among circular vesicles consist of a broadening

of sizes, both larger and smaller (Figure 3.9.A). Although both test PSI:PL concentrations

resulted in several larger circular vesicles, the 16 µM PSI/1000 µM PL incubations resulted

in a significantly lower average diameter compared to both control and the 10 µM PSI/500

µM PL test (P≤0.05). The average diameter (not to be confused with size distribution,

above) for the latter group was not significantly different from control (P≥0.05).

88

Figure 3.9: Summary of phenomena from cryo-TEM images. (A) Enumeration and

comparison by size of vesicle morphologies as identified in the legend; (B) histogram comparison

of circular vesicle size distributions; and (C) histogram comparison of all vesicle morphologies’

size distributions. Note that results are derived from images not limited to the selected examples in

Figure 8, and exclude vesicles in contact with the sample grid.

89

Nevertheless, it is clear that increased diameter vesicles appear in test images irrespective

of diameter averages within a broadened size population.

With respect to imaging, a challenge that was encountered concerned vesicles’ tendency

to preferentially adhere to the sample support grid instead of becoming entrapped within

the pores. As reflected in the relatively low number of vesicle measurements for the control

group in Figure 3.9, vesicle adherence was particularly problematic for non-treated

vesicles. This was even more pronounced for 1:1:1 POPC:POPE:POPS vesicles where

insufficient quality of the control images resulted in exclusion of test images (i.e., virtually

no vesicles were observed within the sample support grid pores while countless adhered to

the grid material). Images for the excluded PSI-POPC:POPE:POPS tests showed similar

morphologies to those described above (data not shown). Further characterizations as part

of a more comprehensive TEM-based study will address the adherence problem.

3.4.7 Characterization of Bilayer Fusion Activity: Dynamic Light Scattering - DLS

was used to measure apparent size changes in 1:1 POPE:POPS LUVs over time as a means

of monitoring fusogenic activity as has been used previously (Bryksa et al, 2011; Michalek

& Leippe, 2015; Yang et al, 2015). Assays were carried out at room temperature until

polydispersity or aggregation of samples became excessive for DLS measurements, up to

26 h. LUV size was monitored in both non-reducing and reducing buffered saline (Figure

3.10.A-B, respectively). Native PSI induced vesicle changes within 5 min, causing peak

broadening. By 30 min, a population of large particles (4000±1000 nm) constituted ~2%

of the vesicle population and the remainder constituted a size range averaging almost

double their original apparent size (220±100 nm). Beyond 60 min, aggregates began to

form thus precluding further DLS measurements. Fusion assays for PSI in reducing

90

Figure 3.10: PSI-treated vesicle size changes over time monitored by DLS. 1:1

POPE:POPS LUVs (100 µM PL) were incubated with 10 µM PSI in either (A) non-reducing or

(B) reducing buffered saline at pH 4.5.

91

conditions also initially shifted the vesicle population to higher average diameter, however,

a broader and larger size population formed more quickly compared to native PSI. This

effect was also reflected in excessively high polydispersity prior to the formation of

aggregates. In the absence of its disulfide bonds, PSI appeared to produce intermediate size

products, shifting the entire population of LUVs, unlike native PSI which tended to

increase the size range (broadening of the initial peak in Figure 3.10) via more modest and

gradual size increases.

3.4.8 PSI Component Peptide-Induced Bilayer Disruption - Peptides corresponding to

the secondary structure subdivisions in PSI (3RFI), as outlined in Figure 3.1, were screened

for bilayer disruption activity by LUV leakage assays at concentrations up to 20 µM using

1:1 POPE:POPS LUVs (100 µM PL) in buffered saline pH 4.5. The results are summarized

in Figure 3.11.A where H3 was the only peptide that induced leakage, significantly

different from each of the other groups including negative control tests (P≤0.05). Like PSI,

H3 leakage activity plateaued at approximately 25% of positive control, however, rates

were slower than the full length protein (Figure 3.5) such that calcein leakage proceeded

for approximately twice as long (75-90 min). Leakage assays were also done for

combinations of peptides, none of which yielded activity beyond that of H3 alone (data not

shown). Although H3 was subsequently confirmed to induce leakage in both reducing and

non-reducing conditions, all peptide leakage data analyzed below were restricted to rates

measured in buffered saline containing 10 mM DTT. This was done to prevent

oligomerization of peptides via unpaired Cys residues which would normally be contained

in one of three disulfide bonds within native PSI (see Figure 3.1).

92

Figure 3.11: Leakage assays for peptide-treated 1:1 POPE:POPS LUVs at 25° C in

buffered saline pH 4.5. (A) Peptide screening assays for LUV leakage activity; (B) Leakage

rate-dependence on [PL]:[H3] ratio; and (C-D) Extent of leakage induced by 4-20 µM H3 (mean

+/- one standard deviation, n=3). Leakage was non-linearly proportional to H3 concentration, and

semi log curves were significantly different (P≤0.05). Error bars indicate +/- one standard

deviation, n=3.

93

H3-induded bilayer disruption was further characterized in terms of concentration

dependence of leakage. First, maximum leakage rate was confirmed to be non-linearly

dependent on [PL]:[H3] ratio (Figure 3.11.B). Second, insight into the mode of H3-induced

bilayer disruption was investigated by relating peptide concentration to the extent of

leakage at 10, 20 and 30 min (Figure 3.11.C-D). The semi-log analyses (Figure 3.11.D)

revealed distinct (P≤0.05) linear curves having increasing slopes (P≤0.05) and similar

intercepts (P˃0.05) (see inset table in Figure 3.11.D). Solving for the x-intercepts, there

appeared to be a critical peptide concentration (average 2.9±0.5 µM) necessary for

leakage/disruption activity.

3.4.9 H3 Intrinsic Tryptophan Fluorescence in the presence of Anionic Bilayer

Vesicles - Since H3 contains one of StAP PSI’s two Trp residues, and H3 shows fusion and

leakage activities, Trp emission was measured upon incubation with 1:1 POPE:POPS

LUVs to determine whether its bilayer interactions involve bilayer penetration by its

hydrophobic mid sequence MAVVWM (see Figure 3.1) . Assays were carried out in the

same manner as for PSI and the resulting spectra are presented in Figure 3.12. Like PSI,

λmax did not shift, and initially, H3 fluorescence emission gradually increased up to the 95

s scan. However, emission subsequently trended down, overshooting the start scan level

and then returned to near the initial 20 s scan. Although changes between consecutive scans

seemed to indicate a coherent progression, scans were all within error of one another,

leading to the conclusion that H3 bilayer penetration is unlikely.

3.4.10 Secondary Structure of PSI Component Peptides - Each of the PSI secondary

structure component peptides (Figure 3.1), as well as a H3 Lys83Gln mutant, were

characterized in terms of secondary structure by far-UV CD (Figure 3.13.A-F). Since the

94

Figure 3.12: Intrinsic Trp fluorescence emission spectra upon incubation of 10 µM

peptide H3 with 1:1:1 POPC:POPE:POPS LUVs (100 µM PL) in buffered saline pH

4.5 at 25° C. Times indicated correspond to scan start times (seconds) after initial mixing.

Although peak emission increased over the initial 95 s, signal subsequently decreased until 225 s

at which point it remained constant over scans collected through 405 s (overlapping scans omitted

for clarity). No shift in λmax was detected. Note that each spectrum is the average of three scans,

error bars were omitted for clarity in distinguishing the spectra, and differences in λmax and

Emissionmax were not significantly different (P>0.05).

95

Figure 3.13: Far-UV CD spectra of peptides in buffered saline containing 10 mM DTT

at the indicated pH values. Spectra were averaged from three scans: (A) H1; (B) H2; (C) H1H2;

(D) X; (E) H3 (solid lines), H3 with 1:1:1 POPC:POPE:POPS LUVs (100 µM PL; broken lines)

and Lys83Gln H3 (dotted lines); and (F) H4.

96

peptides (except X) contain Cys, spectra were collected in reducing conditions. Firstly, in

the absence of the PSI structure, peptides H2, X and H4 (Figure 3.13.B,C,F, respectively)

have secondary structures that are predominantly disordered/random coil (large negative

peak centred at 200 nm) possibly mixed with helix or β-structure sufficient to cause an

overall negative ellipticity across most of the spectra, a region that would have near zero

CD signal if these peptides were entirely disordered (Kelly & Price, 2000; Whitmore &

Wallace, 2008). The high coil character of these spectra was present irrespective of

reducing or non-reducing conditions.

The most notable findings among the peptides’ spectra concerned the PSI N-terminal

helix, H1 (Figure 3.13.A,C), and H3, the peptide having apparent bilayer fusion and

disruption activities (Figure 3.13.E). The spectra for H1 and H1H2 (Figures 3.13.A,C)

indicate a distinct structural transition delineated by the activity pH profile for PSI such

that pH 3.0 and 4.5 have similar spectra, distinct from those at pH 6.2 and 7.4. It is apparent

that as pH transitions to inactive pH, the PSI N-terminus becomes disordered. In terms of

active pH conditions, H1 and H1H2 appear to have substantial β-structure. With respect to

H3, secondary structure content appears to be dominated by β-structure across the four pH

values tested as all four of the spectra contain minima near 218 nm. H3 Lys83Gln, which

eliminated fusion activity, also had an accompanying drastic effect on H3 secondary

structure in that it became largely disordered across all pH values tested (Figure 3.13.E).

3.4.11 Peptide-induced vesicle fusion – Peptides were screened for fusogenic activity

at up to 20 µM. As was the case for leakage activity, only peptide H3 induced fusion

(Figure 3.14). Similar to PSI (Figure 3.10), H3 caused broadening of the original

population comparatively slower (e.g., Peak 1: 154±50 nm vs. 256±120 nm, respectively,

97

Figure 3.14: H3-treated vesicle size monitored over time by DLS. 1:1 POPE:POPS LUVs

(100 µM PL) were incubated in buffered saline pH 4.5 containing 10 mM DTT with either (A) 10

µM H3 or (B) 10 µM Lys83Gln H3.

98

at 60 min). H3 caused an immediate appearance of very large particles by 5 min at an

abundance of 4%, whereas PSI did not produce any large particles as of the 15 min reading.

Since Lys83 provides the sole positive charge within H3, mutant Lys83Gln H3 was also

assayed to test its suspected role in PSI interaction with anionic bilayers. Lys83Gln H3

induced no discernible change in vesicle size through 26 h indicating that Lys83 is essential

for PSI-bilayer interaction.

3.5 Discussion

Dimer formation has been implicated in saposin-membrane interactions in several

structure-function studies (Ahn et al, 2003; Hawkins et al, 2005; John et al, 2006;

Rossmann et al, 2008). Although the crystal structure of StAP PSI (3RFI) revealed a dimer

(Bryksa et al, 2011), PSI quaternary structure in solution remained unclear. Determining

the pH-dependence of PSI dimerization was thus a priority for understanding its

mechanism of action. Sedimentation equilibrium (Figure 3.2) and Trp fluorescence

emission (Figure 3.3) experiments together indicated that PSI exists as a dimer at low pH

and as a monomer at neutral pH. At pH 6.2, there appeared to be a small proportion of

dimer mixed into the predominantly monomer population since the calculated apparent

mass (Figure 3.2) was slightly higher compared to pH 7.4 (P≤0.05). Considering that pH

6.2 is within a unit of endosomal/vacuolar pH in some circumstances (Nozue et al, 1997),

it seems reasonable for a pH-induced structure transition in PSI to begin near this point.

Since reduced solubility precluded analytical centrifugation at pH 4.5, comparison of Trp

fluorescence measurements at four pH values (Figure 3.3) was used to infer the quaternary

structural state. Indistinguishable spectra at pH 3.0 and 4.5 indicated that PSI is a dimer at

99

pH 4.5. In this context, the findings that bilayer disruption was strong at pH 3.0 and 4.5,

negligible at pH 6.2 and inactive at pH 7.4 (Figure 3.5), and that apparent vesicle fusion

occurs exclusively in acidic conditions (Figure 3.10.A), together indicate that the PSI dimer

is the initial protein structural state in the swaposin-bilayer interaction mechanism.

Monitoring Trp fluorescence emission upon mixing PSI with bilayer vesicles (Figure

3.6) indicated a time-dependent emission increase. After 5 min, slowing spectral changes

suggested a system approaching equilibrium, presumably reflective of Trp stabilizing in a

new environment (Figure 3.8.B). Unexpectedly, there was no change in λmax, a

counterintuitive result in consideration of the increasing emission signal for a system

having a tight, very hydrophobic core PSI dimer (Bryksa et al, 2011) as the start structure.

The overall net increase in Trp emission would presumably derive from the development

of an even more hydrophobic environment accompanied change in λmax. Possibly, the lack

of blue shift was caused by phenomena unrelated to PSI Trp residues’ bilayer environment.

An internal Stark effect (Vivian & Callis, 2001) can be manifested if a positively charged

residue becomes positioned near the Trp benzene ring, or a negative charge near the pyrrole

ring, causing a λmax red-shift via a process outside of the intended experimental design. H1

and H3 contain Trp18 and Trp77 (non-recombinant StAP PSI numbering), respectively:

(H1) IVSMECKTIVSQYGEMIWDLLVSG

(H3) EAPLCTACEMAVVWMQNQLKQ

Trp18 (underlined italics) is neighbored by an ionisable residue Glu19 (underlined) that

could reasonably undergo a change in charge induced by an altered pKa upon protein

structure change and/or contact or even penetration of PL bilayer. Further contributing

charge in the vicinity in question is the negatively charged PL bilayer itself, again

100

potentially directly impacting the electric field near the Trp pyrrole ring (Vivian & Callis,

2001). In addition to these factors, another potentially confounding situation could be the

formation of oligomers such that Trp becomes buried in a hydrophobic protein environment

(Chen & Barkley, 1998; Shai, 1999) other than its dimer start structure.

Insights into PSI-induced effects on bilayer vesicle size and morphologies were gained

from analysis of DLS and cryo-TEM data. DLS experiments showed that PSI caused

broadening of vesicle size range and shifting to larger vesicle diameters (Figure 3.10).

Cryo-TEM images revealed several different structures (Figure 3.8.B-C) whose

characteristics are summarized in Figure 3.9. Anionic vesicle interaction with sample grids

and limited access constrained our choice of concentration and lipid:protein ratio

combinations. Cryo-TEM imaging of 15-minute vesicle incubations with PSI showed

structures that spanned a significantly broadened size distribution compared to untreated

control vesicles (see Figure 3.9). Comparison to the DLS size distribution at 15 min

suggests overall agreement on moderately broadened size ranges, the appearance of both

smaller and larger vesicles compared to controls, and an absence of particles exceeding

~600 nm. Furthermore, the appearance of individual spheroid vesicles having larger

diameters than control samples in cryo-TEM images, taken together with the DLS and

TEM size distributions’ similarities, give direct evidence of PSI-mediated vesicle fusion.

The varied vesicle sizes and morphologies detailed in Figure 3.9 may represent different

stages of vesicle collapse when viewed in the context of TEM image interpretation

discussed previously (Almgren et al, 2000; Edwards et al, 1993). The array of shapes

appears to follow a progression from initial flattening/creation of a liposome edge (pink

arrows), followed by expansion of edging such that wedge type shapes may form (yellow

101

arrows), followed by elongated narrow vesicles (blue arrows), and finally a fully flattened

dense elongated structure (red arrows). The semi-circular single edge morphology is a

known phenomenon regarding vesicle-fusase initial interactions (Gui et al, 2016), and it

has also been reported for melittin-induced vesicle effects (Strömstedt et al, 2007).

Regarding secondary structure, the present study confirms the pH-activity-helix

structure increase relationship not only by the solution CD determinations, but also via

evidence produced from the PSI-vesicle CD analyses (Figure 3.7) and intrinsic Trp

fluorescence measurements (Figure 3.6). Qualitatively, spectra for PSI with LUVs and

NaCl indicated increased ellipticity peak magnitudes at 222 and 208 nm, and were

distinctive and consistent at the two concentrations of PSI tested. This structure change is

solely dependent upon interaction with PL bilayer as the PSI samples themselves had

already been equilibrated in CD buffer prior to scans. These findings lead to the conclusion

that PSI-PL bilayer induces secondary structure changes resulting in an increase in helix

structure.

Furthermore, it is worth noting that native and reducing conditions yielded very similar

CD spectra for respective pH values. Disulfide bonds, it would appear, have little role in

the overall secondary structure content of the PSI in saline conditions. PSI DLS profiles

(Figure 3.10) for native and reducing conditions also indicated that disulfide bonds were

non-essential for vesicle fusion. In contrast to vesicle size range broadening and increasing

diameters by native PSI, elimination of PSI disulfide bonds in reducing conditions uniquely

resulted in a relatively rapid and complete shift to discreet populations of larger vesicles.

We reasoned that a disulfide bond-free PSI structure would allow increased accessibility

of protein structural features, perhaps resulting in the altered vesicle size profiles in Figure

102

3.10. In this context, altering overall PSI structure appears to change the kinds of fusion

products, but does not prevent it from occurring. It would thus appear that at least basic

PSI-bilayer interaction must be embedded in sequence(s) or structural feature(s)

independent of quaternary and tertiary structure considerations.

Secondary structure comparisons over different pH conditions indicated important

relationships between structure and environmental changes not only for PSI, but also

among its component helices. The clear trend of increasing helix and decreasing strand

structure content upon acidification was expected based on previous work (Bryksa et al,

2011) which only considered pH 7.4 and 4.5. With respect to peptide secondary structures,

it was hypothesized that the pH-induced increase in helix content would originate from

some portion of the overall PSI structure. The reasoning for this expectation was rooted in

that the PSI is a domain-swapped saposin in terms of the arrangement of the N- and C-

terminal halves of the respective monomers. That is, the PSI N- and C-terminal halves

occur in the opposite order relative to saposins (i.e., H1-H2-H3-H4 becomes H3-H4-H2-

H1). Despite this drastic difference at the tertiary structure level, PSI and saposin C (and

to a lesser extent saposin D) share highly similar pH dependencies as well as anionic PL

requirement for activity (Ahn et al, 2003; Ahn et al, 2006; Alattia et al, 2006; Ciaffoni et

al, 2001; Ciaffoni et al, 2006; de Alba et al, 2003; Hawkins et al, 2005; Mendieta et al,

2006; Vaccaro et al, 1995a; Vaccaro et al, 1995b; Vaccaro et al, 1994), and PSI and

saposins B and C share highly similar tertiary structures (Ahn et al, 2003; Bryksa et al,

2011; Rossmann et al, 2008). Hence, we hypothesized that the structure-function

“controls” must be intrinsic features originating from a lower structural level than the

backwards PSI “swaposin” tertiary structure.

103

The activities of the component peptides, measured in reducing conditions to prevent

inter- and intramolecular cross-linking, produced some unexpected results. The bilayer

disruption activity of H3 (Figure 3.11) disproved the hypothesis that the C-terminus of PSI

(equivalent to the N-terminus of saposins) is the source of PSI bilayer disruption.

Furthermore, H3 caused bilayer fusion in PL vesicles (Figure 3.14) while none of the other

peptides resulted in leakage or fusion. The fusion profiles for H3 and PSI are similar

overall, however, H3 caused apparent fusion faster and was seemingly more efficient in

creating a higher proportion of LUVs having large (~4000 nm) particle sizes. Component

regions of saposin C have been studied previously (Wang et al, 2003) with notably different

results. For saposin C, H1 and H2 (equivalent to PSI H3 and H4, respectively) did not cause

bilayer fusion. Comparing sequences, PSI H3 is more hydrophobic than either of the

saposin C peptides, and although the positive charge content is inferior for PSI H3, its

amphipathic charge distribution, together with the hydrophobic AVVW middle patch, may

explain the different fusogenic abilities for the respective regions of these two otherwise

similar proteins:

PSI H3 - EAPLCTACEMAVVWMQNQLKQ

SapC H1 - YCEVCEFLVKEVTKLID

SapC H2 - EKEILDAFDKMCSKLPK

Upon finding that H3 had both leakage and fusion activities, three mutants were

designed each eliminating the respective charges: Glu63Gln, Glu74Gln and Lys83Gln.

While the sole Lys was predicted to be essential for anionic bilayer interaction, the other

two charges were suspected of potentially having a role in governing pH-dependence of

structure. Unfortunately, low solubility for both Glu mutants prevented their use, at least

under the conditions used in the present study. H3 Lys83Gln was soluble and was shown

104

to lack both bilayer disruption and vesicle fusion activities, thus confirming that Lys83 is

critical for H3-bilayer interaction. As a result of this finding, its role is being investigated

as part of a larger mutational study on PSI charge features.

From analysis of H3 bilayer disruption activity (Figure 3.11), the apparent requirement

for attaining a critical peptide concentration to initiate leakage suggests a cooperative

mechanism of action. The findings of the present study regarding H3 share certain

characteristics with other membrane-active peptide types, including magainins (Matsuzaki

et al, 1989; Matsuzaki et al, 1996; Matsuzaki et al, 1998) and amyloid peptides (Burke et

al, 2013; Butterfield & Lashuel, 2010). In common with the magainins is that they yield

abnormally slow rates of vesicle leakage (Matsuzaki et al, 1991; Matsuzaki et al, 1989),

that leakage increases dramatically with peptide concentration once the concentration

threshold is met, that leakage rate slows with time for all peptide concentrations (Matsuzaki

et al, 1989; Vandenburg et al, 2002), and that leakage stops prior to attaining full leakage

(Matsuzaki et al, 1991). The above kinetic features, all in common with the findings of the

present study, indicate that H3 may operate by a carpet-like mechanism (Vandenburg et al,

2002) wherein a cooperative bilayer perturbing process results in leakage of vesicle

contents, functionality which subsequently dissipates due to peptide equilibration across

the bilayer (Matsuzaki et al, 1989; Vandenburg et al, 2002). In the case of amyloid-β

peptide, calcein leakage from peptide-treated LUVs also proceeds only to partial leakage

(McLaurin & Chakrabartty, 1997; Williams et al, 2010). More importantly perhaps, certain

amyloid peptides, proposed to employ the carpet mechanism in some situations (Engel et

al, 2008; Hebda & Miranker, 2009), produce similar distortions in anionic LUV shape (i.e.,

the appearance of flat surfaces and elongated vesicles) to those of the present study

105

observed by cryo-TEM (Engel et al, 2008; Williams et al, 2010; Yip & McLaurin, 2001).

The implications of these similarities remain to be elucidated in future investigations.

Although unexpected, the possibility of shared mechanistic bilayer disruption features

with amyloid peptides seems reasonable in the context of the predominantly β-strand/turn

secondary structure (Figure 3.13.E). While it was found that PSI undergoes a secondary

structure transition such that helix is gained at the expense of strand upon acidification

(Figure 3.4.D) and concomitant dimerization (Figure 3.2), and this gain in helix modestly

increases upon interaction with bilayer (Figure 3.7), H3 alone seemingly has no substantial

helix content nor does its structure appear to transition to helix upon bilayer interaction or

have any obvious pH-sensitivity at all (Figure 3.13). These key differences indicate that

PSI and H3 may have two different mechanisms of action, a possibility that is further

supported by the different trends in Trp emission spectral changes upon mixing with bilayer

vesicles; PSI Trp gain hydrophobicity reaching a new and stable state (Figure 3.6) whereas

H3 appeared to experience short-lived changes in hydrophobicity only to return to the start

spectrum within approximately 5 min and remaining static thereafter (Figure 3.12). The

lack an increase in H3 Trp hydrophobicity indicates a non-penetrating mechanism of

action, consistent with the aforementioned carpet model wherein peptides interact

electrostatically with PL headgroups covering regions of the bilayer surface (Bobone,

2014). By contrast, increased PSI Trp hydrophobicity reaching equilibrium within ~5 min

of encountering bilayer, with a half time of ~30 s (Figure 3.6.B), agreed with the

approximate time required for detectable secondary structure changes to reach their

apparent end point (Figure 3.7); where obvious spectral shifting was underway by 30 s and

essentially completed by ~4.5 min. The increased hydrophobicity suggests insertion into

106

the bilayer such that Trp interact with PL tail moieties. These features, together with the

finding of a critical [H3] for commencing leakage, are consistent with the previously

proposed mechanism of StAP PSI bilayer disruption as involving pore formation as per the

Barrel-Stave model (Muñoz et al, 2014), which consists of pore formers oligomerizing with

hydrophobic moieties facing away from the pore opening and interacting with hydrophobic

inner bilayer (Phoenix et al, 2013). This model may also account for incomplete leakage

noted in the present study in that pore formers may be situated in a fraction of vesicles

thereby causing leakage in just some targets (i.e., all-or-none leakage) while others remain

intact (Last et al, 2013).

H3, while apparently membrane-active irrespective of PSI, is particularly interesting in

regards to its sequence features. H3 has desirable elements characteristic of particular

antimicrobial sequences:

EAPLCTACEMAVVWMQNQLKQ

First, there is a single positive charge (bold underlined) that was confirmed in the present

study to interact with anionic PL bilayer; second, the peptide is relatively small at 21

residues yet is amphiphilic (Glu63 and Glu74 underlined; and Lys83 bold underlined);

third, it contains a hydrophobic four-residue patch centred within the sequence (underlined

italics); and fourth, it contains multiple Gln residues in the vicinity of a positive charge

(bold), common for antimicrobial cationic peptides (Patrzykat & Douglas, 2005), although

typically in conjunction with more positive charge than present in H3 (Sharifahmadian et

al, 2013). The C-terminus of H3 corresponds to a membrane-associated protein motif

107

typically found in the plant defense-related 2S storage proteins and chitin binding proteins,

in addition to some saposins:

[N/Q]-[N/Q]-[A/L/I/V]-[K/R]-[N/Q].

This motif is present in many flocculating/coagulating proteins (see Table 3.1 for

examples) where Asn and Gln, and Lys and Arg are interchangeable, respectively, and the

third position from the C-terminus can be any aliphatic residue. Recently, bacterial

membrane damage by Moringa oleifera seed flocculating cationic peptide (see Table 3.1)

was shown to be caused by a membrane fusion mechanism (Shebek et al, 2015). Also, it

was reported recently that the antibacterial activity of a Ricinus communis 2S albumin

involves leakage of cell contents postulated to be the result of magainin-like (noted above)

pore formation (Souza et al, 2016). A Motif Query of the RCSB Protein Data Bank

(avoiding redundancy) for the full C-terminal 6-residue motif [N/Q]-[N/Q]-[N/Q]-

[A/L/I/V]-[K/R]-[N/Q] yielded just 28 PDB structures in total (Table 3.2), all of which are

lipid-interacting or surface active proteins.

A structure alignment search of the RCSB PDB structure database based on protein

domains (as opposed to whole protein chains) yielded matches for diverse surface-

active/membrane-interacting proteins. High-scoring alignments are presented in Figure

3.15. Interestingly, among these are endosomal/vacuolar factors (Figure 3.15.A,C), a seed

storage albumin protein (Figure 3.15.F) and a calcium sensing protein (Figure 3.15.E). The

latter is particularly noteworthy due to its utilization of the calcium binding protein EF-

hand motif, i.e., a helix-loop-helix segment usually occurring in adjacent pairs (Lewit-

Bentley & Réty, 2000), reminiscent of a similar N-terminal kinked helix motif within the

side-by-side boomerang-shaped open saposin fold of PSI (Bryksa et al, 2011).

108

Table 3.1: Examples of antimicrobial/membrane-interacting proteins that contain the

[N/Q]-[N/Q]-[A/L/I/V]-[R/K]-[N/Q] motif. Residues/sequences are indicated as:

hydrophobic (bold black underlined), the motif in question (bold red underlined) or similar

sequence to that motif (bold red).

Protein Sequence Ref.

StAP PSI H3 EAPLCTACEMAVVWMQNQLKQ (Guevara et al.,

2005)

Mabinlin I

EPLCRRQFQQHQHLRACQRYIRRRAQ

RGGLVDEQRGPALRLCCNQLRQVNKP

CVCPVLRQAAHQQLYQGQIEGPRQVR

QLFRAARNLPNICKIPAVGRCQFTRW

(Nirasawa et al.,

1994)

Mabinlin II

QPRRPALRQCCNQLRQVDRPCVCPVL

RQAAQQVLQRQIIQGPQQLRRLFDAA

RNLPNICNIPNIGACPFRAW

(Li et al., 2008; Liu

et al., 1993)

Flocculent-active

protein MO2.1

and MO2.2

QGPGRQPDFQRCGQQLRNISPPQRCPS

LRQAVQLTHQQQGQVGPQQVRQMYR

VASNIPST

(Broin et al., 2002)

Moringa oleifera

CBP3 CPAIQRCCQQLRNIQPPCRCCQ

(Gifoni et al.,

2012)

Sesame 2S

albumin

(Sesamum

indicum)

MAKKLALAAVLLVAMVALASATTYT

TTVTTTAIDDEANQQSQQCRQQLQGR

QFRSCQRYLSQGRSPYGGEEDEVLEMS

TGNQQSEQSLRDCCQQLRNVDERCRC

EAIRQAVRQQQQEGGYQEGQSQQVY

QRARDLPRRCNMRPQQCQFRVIFV

(Tai et al., 2001)

* β-lactoglobulin

fragment 1−8

LIVTQTMK * This example is shown because it contains the

skeleton of the motif in question; hydrophobic

patch, Gln, single positive charge

(Pouliot et al.,

2009)

109

Table 3.2: Proteins in the RCSB PDB databank that contain the [N/Q]-[N/Q]-[N/Q]-

[A/L/I/V]-[R/K]-[N/Q] motif.

110

Figure 3.15: Protein domain structure alignments. A structure alignment search of the

RCSB PDB database using the Protein Structure Comparison Tool v.4.2.0 (Prlić et al., 2010)

running the FATCAT algorithm (Ye and Godzik, 2003) based on protein domains (as opposed to

whole protein chains) yielded matches to PSI (3RFI) for diverse surface active/membrane

interacting proteins. The alignments presented here were selections based upon a combination of

high alignment score, low P-value and/or high sequence similarity: (A) Endosomal sorting

complex, (B) sodium channel protein type 5 subunit α, (C) vacuolar transport chaperone 4, (D)

calmodulin, (E) jellyfish calcium-regulated photoprotein, and (F) sweet protein mabinlin-2.

111

In conclusion, inspection of the structural elements that make up StAP PSI has offered

insights into its membrane interaction requirements, and has offered more questions for

future consideration. The present study showed that, nested within its primary structure,

PSI contains a peptide (H3) that is itself bilayer-active. Oddly, this peptide takes up a β-

stand configuration when alone as opposed to its normal helix character within the PSI

tertiary structure. H3 deserves further investigation to assess its biological activity and

mutability as it may present a novel sequence for membrane active protein applications

(e.g., nano-particle targeting), irrespective of its apparent central role in PSI-bilayer

structure-function. The H1 region, part of a very hydrophobic region of PSI that tightly

binds within the dimer (Bryksa et al, 2011), was shown to be the sole portion of PSI that is

intrinsically pH-sensitive albeit adopting a strand configuration at acidic pH and is

otherwise unstructured, as is the case at all pH values for H2, “X” and H4. Thus, the overall

tertiary and quaternary structures of PSI are essential to the folding of every component

native helix, including both the dimer and monomer PSI states. Ironically, elimination of

PSI’s disulfide bonds, which ubiquitously hold the N- and C-terminal portions of SAPLIPs

together, does not result in appreciable loss of native secondary structure. This leaves one

evident source of PSI-induced helix: the extensive hydrophobic contacts that reside within

both the extended dimer as well as the canonical monomeric saposin fold. At present,

investigations into the PSI transition to the bilayer environment as well as the roles of

charged residues seek to better illuminate the structural basis for PSI bilayer interactions.

More broadly, the apparent structural cohesion among surface/lipid active proteins across

protein families is a vast topic that calls for detailed comparative structural investigation in

the context of plant immunity as well as pathogen invasion factors.

112

3.6 References

Ahn, V. E., Faull, K. F., Whitelegge, J. P., Fluharty, A. L. & Prive, G. G. (2003) Crystal

structure of saposin B reveals a dimeric shell for lipid binding. Proceedings of the

National Academy of Sciences of the United States of America, 100(1), 38-43.

Ahn, V. E., Leyko, P., Alattia, J.-R., Chen, L. & Privé, G. G. (2006) Crystal structures of

saposins A and C. Protein Science, 15(8), 1849-1857.

Alattia, J.-R., Shaw, J. E., Yip, C. M. & Privé, G. G. (2007) Molecular imaging of

membrane interfaces reveals mode of β-glucosidase activation by saposin C.

Proceedings of the National Academy of Sciences, 104(44), 17394-17399.

Alattia, J. R., Shaw, J. E., Yip, C. M. & Prive, G. G. (2006) Direct visualization of saposin

remodelling of lipid bilayers. Journal of Molecular Biology, 362(5), 943-953.

Almgren, M., Edwards, K. & Karlsson, G. (2000) Cryo transmission electron microscopy

of liposomes and related structures. Colloids and Surfaces A: Physicochemical and

Engineering Aspects, 174(1–2), 3-21.

Anderson, D. H., Sawaya, M. R., Cascio, D., Ernst, W., Modlin, R., Krensky, A. &

Eisenberg, D. (2003) Granulysin crystal structure and a structure-derived lytic

mechanism. Journal of Molecular Biology, 325(2), 355-365.

Andreu, D., Carre, C., Linde, C., Boman, H. G. & Andersson, M. (1999) Identification of

an anti-mycobacterial domain in NK-lysin and granulysin. Biochemical Journal, 344(3),

845-849.

Bobone, S. (2014) Peptide and Protein Interaction with Membrane Systems: Applications

to Antimicrobial Therapy and Protein Drug Delivery. Switzerland: Springer.

Broin, M., Santaella, C., Cuine, S., Kokou, K., Peltier, G. & Joët, T. (2002) Flocculent

activity of a recombinant protein from Moringa oleifera Lam. seeds. Applied

Microbiology and Biotechnology, 60(1), 114-119.

Bruhn, H. (2005) A short guided tour through functional and structural features of saposin-

like proteins. Biochemical Journal, 389, 249-257.

Bryksa, B. C., Bhaumik, P., Magracheva, E., DeMoura, D. C., Kurylowicz, M., Zdanov,

A., Dutcher, J. R., Wlodawer, A. & Yada, R. Y. (2011) Structure and mechanism of the

113

saposin-like domain of a plant aspartic protease. Journal of Biological Chemistry,

286(32), 28265-28275.

Burke, K. A., Yates, E. A. & Legleiter, J. (2013) Biophysical insights into how surfaces,

including lipid membranes, modulate protein aggregation related to neurodegeneration.

Frontiers in Neurology, 4, 17.

Butterfield, S. M. & Lashuel, H. A. (2010) Amyloidogenic protein–membrane interactions:

Mechanistic insight from model systems. Angewandte Chemie International Edition,

49(33), 5628-5654.

Chen, Y. & Barkley, M. D. (1998) Toward understanding tryptophan fluorescence in

proteins. Biochemistry, 37(28), 9976-9982.

Ciaffoni, F., Salvioli, R., Tatti, M., Arancia, G., Crateri, P. & Vaccaro, A. M. (2001)

Saposin D solubilizes anionic phospholipid-containing membranes. Journal of

Biological Chemistry, 276(34), 31583-31589.

Ciaffoni, F., Tatti, M., Boe, A., Salvioli, R., Fluharty, A., Sonnino, S. & Vaccaro, A. M.

(2006) Saposin B binds and transfers phospholipids. Journal of Lipid Research, 47(5),

1045-1053.

Curto, P., Lufrano, D., Pinto, C., Custódio, V., Gomes, A. C., Trejo, S. A., Bakás, L., Vairo-

Cavalli, S., Faro, C. & Simões, I. (2014) Establishing the yeast Kluyveromyces lactis as

an expression host for production of the saposin-like domain (plant-specific insert) from

the aspartic protease cirsin. Applied and Environmental Microbiology, 80(1), 86-96.

de Alba, E., Weiler, S. & Tjandra, N. (2003) Solution structure of human saposin C: pH-

dependent interaction with phospholipid vesicles. Biochemistry, 42(50), 14729-14740.

Edwards, K., Gustafsson, J., Almgren, M. & Karlsson, G. (1993) Solubilization of lecithin

vesicles by a cationic surfactant: intermediate structures in the vesicle-micelle transition

observed by cryo-transmission electron microscopy. Journal of Colloid and Interface

Science, 161(2), 299-309.

Egas, C., Lavoura, N., Resende, R., Brito, R. M., Pires, E., de Lima, M. C. & Faro, C.

(2000) The saposin-like domain of the plant aspartic proteinase precursor is a potent

inducer of vesicle leakage. Journal of Biological Chemistry, 275(49), 38190-38196.

Engel, M. F. M., Khemtémourian, L., Kleijer, C. C., Meeldijk, H. J. D., Jacobs, J., Verkleij,

A. J., de Kruijff, B., Killian, J. A. & Höppener, J. W. M. (2008) Membrane damage by

114

human islet amyloid polypeptide through fibril growth at the membrane. Proceedings

of the National Academy of Sciences, 105(16), 6033-6038.

Frazão, C., Bento, I., Costa, J., Soares, C. M., Veríssimo, P., Faro, C., Pires, E., Cooper, J.

& Carrondo, M. A. (1999) Crystal structure of cardosin A, a glycosylated and Arg-Gly-

Asp-containing aspartic proteinase from the flowers of Cynara cardunculus L. Journal

of Biological Chemistry, 274(39), 27694-27701.

Ghannoum, M. A. & Rice, L. B. (1999) Antifungal agents: Mode of action, mechanisms

of resistance, and correlation of these mechanisms with bacterial resistance. Clinical

Microbiology Reviews, 12(4), 501-517.

Gifoni, J. M., Oliveira, J. T. A., Oliveira, H. D., Batista, A. B., Pereira, M. L., Gomes, A.

S., Oliveira, H. P., Grangeiro, T. B. & Vasconcelos, I. M. (2012) A novel chitin-binding

protein from Moringa oleifera seed with potential for plant disease control. Peptide

Science, 98(4), 406-415.

Glathe, S., Kervinen, J., Nimtz, M., Li, G. H., Tobin, G. J., Copeland, T. D., Ashford, D.

A., Wlodawer, A. & Costa, J. (1998) Transport and activation of the vacuolar aspartic

proteinase phytepsin in barley (Hordeum vulgare L.). Journal of Biological Chemistry,

273(47), 31230-31236.

Goodwin, J. & Khant, H. Frozen hydrated lipid vesicles. http://www.gatan.com-

/products/tem-specimen-preparation/cryoplunge-3-system#publication-related:

National Center for Macromolecular Imaging, Baylor College of Medicine, Houston,

TX, USA.

Guevara, M. G., Oliva, C. R., Huarte, M. & Daleo, G. R. (2002) An aspartic protease with

antimicrobial activity is induced after infection and wounding in intercellular fluids of

potato tubers. European Journal of Plant Pathology, 108(2), 131-137.

Guevara, M. G., Almeida, C., Mendieta, J. R., Faro, C. J., Verissimo, P., Pires, E. V. &

Daleo, G. R. (2005) Molecular cloning of a potato leaf cDNA encoding an aspartic

protease (StAsp) and its expression after P. infestans infection. Plant Physiology and

Biochemistry, 43(9), 882-889.

Gui, L., Ebner, J. L., Mileant, A., Williams, J. A. & Lee, K. K. (2016) Visualization and

sequencing of membrane remodeling leading to influenza virus fusion. Journal of

Virology, 90(15), 6948-6962.

115

Hawkins, C. A., de Alba, E. & Tjandra, N. (2005) Solution structure of human saposin C

in a detergent environment. Journal of Molecular Biology, 346(5), 1381-1392.

Hebda, J. A. & Miranker, A. D. (2009) The Interplay of Catalysis and Toxicity by Amyloid

Intermediates on Lipid Bilayers: Insights from Type II Diabetes. Annual Review of

Biophysics, 38(1), 125-152.

John, M., Wendeler, M., Heller, M., Sandhoff, K. & Kessler, H. (2006) Characterization

of human saposins by NMR spectroscopy. Biochemistry, 45(16), 5206-5216.

Kelly, S. M. & Price, N. C. (2000) The use of circular dichroism in the investigation of

protein structure and function. Current Protein and Peptide Science, 1(4), 349-384.

Kendall, D. A. & MacDonald, R. C. (1982) A fluorescence assay to monitor vesicle fusion

and lysis. Journal of Biological Chemistry, 257(23), 13892-13895.

Kervinen, J., Tobin, G. J., Costa, J., Waugh, D. S., Wlodawer, A. & Zdanov, A. (1999)

Crystal structure of plant aspartic proteinase prophytepsin: inactivation and vacuolar

targeting. EMBO Journal, 18(14), 3947-3955.

Kolter, T. & Sandhoff, K. (2005) Principles of lysosomal membrane digestion: Stimulation

of sphingolipid degradation by sphingolipid activator proteins and anionic lysosomal

lipids. Annual Review of Cell and Developmental Biology, 21(1), 81-103.

Last, N. B., Schlamadinger, D. E. & Miranker, A. D. (2013) A common landscape for

membrane-active peptides. Protein Science, 22(7), 870-882.

Lewit-Bentley, A. & Réty, S. (2000) EF-hand calcium-binding proteins. Current Opinion

in Structural Biology, 10(6), 637-643.

Li, D.-F., Jiang, P., Zhu, D.-Y., Hu, Y., Max, M. & Wang, D.-C. (2008) Crystal structure

of Mabinlin II: A novel structural type of sweet proteins and the main structural basis

for its sweetness. Journal of Structural Biology, 162(1), 50-62.

Liepinsh, E., Andersson, M., Ruysschaert, J. M. & Otting, G. (1997) Saposin fold revealed

by the NMR structure of NK-lysin. Nature Structural Biology, 4(10), 793-795.

Liu, X., Maeda, S., Hu, Z., Aiuchi, T., Nakaya, K. & Kurihara, Y. (1993) Purification,

complete amino acid sequence and structural characterization of the heat‐stable sweet

protein, mabinlin II. European Journal of Biochemistry, 211(1‐2), 281-287.

Lufrano, D., Faro, R., Castanheira, P., Parisi, G., Veríssimo, P., Vairo-Cavalli, S., Simões,

I. & Faro, C. (2012) Molecular cloning and characterization of procirsin, an active

116

aspartic protease precursor from Cirsium vulgare (Asteraceae). Phytochemistry, 81, 7-

18.

Matsuda, J., Vanier, M. T., Saito, Y., Tohyama, J., Suzuki, K. & Suzuki, K. (2001) A

mutation in the saposin A domain of the sphingolipid activator protein (prosaposin) gene

results in a late-onset, chronic form of globoid cell leukodystrophy in the mouse. Human

Molecular Genetics, 10(11), 1191-1199.

Matsuzaki, K., Harada, M., Funakoshi, S., Fujii, N. & Miyajima, K. (1991)

Physicochemical determinants for the interactions of magainins 1 and 2 with acidic lipid

bilayers. Biochimica et Biophysica Acta - Biomembranes, 1063(1), 162-170.

Matsuzaki, K., Harada, M., Handa, T., Funakoshi, S., Fujii, N., Yajima, H. & Miyajima,

K. (1989) Magainin 1-induced leakage of entrapped calcein out of negatively-charged

lipid vesicles. Biochimica et Biophysica Acta - Biomembranes, 981(1), 130-134.

Matsuzaki, K., Murase, O., Fujii, N. & Miyajima, K. (1996) An Antimicrobial Peptide,

Magainin 2, Induced Rapid Flip-Flop of Phospholipids Coupled with Pore Formation

and Peptide Translocation. Biochemistry, 35(35), 11361-11368.

Matsuzaki, K., Sugishita, K.-i., Ishibe, N., Ueha, M., Nakata, S., Miyajima, K. & Epand,

R. M. (1998) Relationship of Membrane Curvature to the Formation of Pores by

Magainin 2. Biochemistry, 37(34), 11856-11863.

McLaurin, J. & Chakrabartty, A. (1997) Characterization of the Interactions of Alzheimer

β-Amyloid Peptides with Phospholipid Membranes. European Journal of Biochemistry,

245(2), 355-363.

Mendieta, J. R., Pagano, M. R., Munoz, F. F., Daleo, G. R. & Guevara, M. G. (2006)

Antimicrobial activity of potato aspartic proteases (StAPs) involves membrane

permeabilization. Microbiology, 152, 2039-2047.

Meyer, V. (2008) A small protein that fights fungi: AFP as a new promising antifungal

agent of biotechnological value. Applied Microbiology and Biotechnology, 78(1), 17-

28.

Michalek, M. & Leippe, M. (2015) Mechanistic insights into the lipid interaction of an

ancient saposin-like protein. Biochemistry, 54(9), 1778-1786.

117

Miteva, M., Andersson, M., Karshikoff, A. & Otting, G. (1999) Molecular electroporation:

a unifying concept for the description of membrane pore formation by antibacterial

peptides, exemplified with NK-lysin. FEBS Letters, 462(1), 155-158.

Montesinos, E. & Bardají, E. (2008) Synthetic antimicrobial peptides as agricultural

pesticides for plant-disease control. Chemistry & Biodiversity, 5(7), 1225-1237.

Mutlu, A. & Gal, S. (1999) Plant aspartic proteinases: Enzymes on the way to a function.

Physiologia Plantarum, 105(3), 569-576.

Muñoz, F., Palomares-Jerez, M. F., Daleo, G., Villalaín, J. & Guevara, M. G. (2011)

Cholesterol and membrane phospholipid compositions modulate the leakage capacity

of the swaposin domain from a potato aspartic protease (StAsp-PSI). Biochimica et

Biophysica Acta - Molecular and Cell Biology of Lipids, 1811(12), 1038-1044.

Muñoz, F., Palomares-Jerez, M. F., Daleo, G., Villalaín, J. & Guevara, M. G. (2014)

Possible mechanism of structural transformations induced by StAsp-PSI in lipid

membranes. Biochimica et Biophysica Acta - Biomembranes, 1838(1, Part B), 339-347.

Muñoz, F. F., Mendieta, J. R., Pagano, M. R., Paggi, R. A., Daleo, G. R. & Guevara, M.

G. (2010) The swaposin-like domain of potato aspartic protease (StAsp-PSI) exerts

antimicrobial activity on plant and human pathogens. Peptides, 31(5), 777-785.

Nirasawa, S., Nishino, T., Katahira, M., Uesugi, S., Hu, Z. & Kurihara, Y. (1994)

Structures of heat-stable and unstable homologues of the sweet protein mabinlin.

European Journal of Biochemistry, 223(3), 989-995.

Nozue, M., Yamada, K., Nakamura, T., Kubo, H., Kondo, M. & Nishimura, M. (1997)

Expression of a vacuolar protein (VP24) in anthocyanin-producing cells of sweet potato

in suspension culture. Plant Physiology, 115(3), 1065-1072.

Pagano, M. R., Mendieta, J. R., Muñoz, F. F., Daleo, G. R. & Guevara, M. G. (2007) Roles

of glycosylation on the antifungal activity and apoplast accumulation of StAPs

(Solanum tuberosum aspartic proteases). International Journal of Biological

Macromolecules, 41(5), 512-520.

Patrzykat, A. & Douglas, S. E. (2005) Antimicrobial peptides: Cooperative approaches to

protection. Protein and Peptide Letters, 12(1), 19-25.

118

Payie, K. G., Weadge, J. T., Tanaka, T. & Yada, R. Y. (2000) Purification, N-terminal

sequencing and partial characterization of a novel aspartic proteinase from the leaves of

Medicago sativa L. (alfalfa). Biotechnology Letters, 22(19), 1515-1520.

Pencer, J. & Hallett, F. R. (2003) Effects of vesicle size and shape on static and dynamic

light scattering measurements. Langmuir, 19(18), 7488-7497.

Pereira, C., Pereira, S., Satiat-Jeunemaitre, B. & Pissarra, J. (2013) Cardosin A contains

two vacuolar sorting signals using different vacuolar routes in tobacco epidermal cells.

The Plant Journal, 76(1), 87-100.

Phoenix, D. A., Dennison, S. R. & Harris, F. (2013) Models for the membrane interactions

of antimicrobial peptides. Antimicrobial Peptides, 145-180.

Pouliot, Y., Guy, M.-M., Tremblay, M., Gaonac’h, A.-C., Chay Pak Ting, B. P., Gauthier,

S. F. & Voyer, N. (2009) Isolation and characterization of an aggregating peptide from

a tryptic hydrolysate of whey proteins. Journal of Agricultural and Food Chemistry,

57(9), 3760-3764.

Prlić, A., Bliven, S., Rose, P. W., Bluhm, W. F., Bizon, C., Godzik, A. & Bourne, P. E.

(2010) Pre-calculated protein structure alignments at the RCSB PDB website.

Bioinformatics, 26(23), 2983-2985.

Qi, X. & Grabowski, G. A. (2001) Differential membrane interactions of saposins A and

C. Journal of Biological Chemistry, 276(29), 27010-27017.

Ramalho-Santos, M., Veríssimo, P., Cortes, L., Samyn, B., Van Beeumen, J., Pires, E. &

Faro, C. (1998) Identification and proteolytic processing of procardosin A. European

Journal of Biochemistry, 255(1), 133-138.

Rossmann, M., Schultz-Heienbrok, R., Behlke, J., Remmel, N., Alings, C., Sandhoff, K.,

Saenger, W. & Maier, T. (2008) Crystal structures of human saposins C and D:

Implications for lipid recognition and membrane interactions. Structure, 16(5), 809-817.

Runeberg‐Roos, P., Tormakangas, K. & Östman, A. (1991) Primary structure of a barley‐

grain aspartic proteinase. European Journal of Biochemistry, 202(3), 1021-1027.

Schaller, A. & Ryan, C. A. (1996) Molecular cloning of a tomato leaf cDNA encoding an

aspartic protease, a systemic wound response protein. Plant Molecular Biology, 31(5),

1073-1077.

119

Shai, Y. (1999) Mechanism of the binding, insertion and destabilization of phospholipid

bilayer membranes by α-helical antimicrobial and cell non-selective membrane-lytic

peptides. Biochimica et Biophysica Acta - Biomembranes, 1462(1–2), 55-70.

Sharifahmadian, M., Sarker, M., Palleboina, D., Waring, A. J., Walther, F. J., Morrow, M.

R. & Booth, V. (2013) Role of the N-terminal seven residues of surfactant protein B

(SP-B). PLoS ONE, 8(9), e72821.

Shebek, K., Schantz, A. B., Sines, I., Lauser, K., Velegol, S. & Kumar, M. (2015) The

flocculating cationic polypetide from Moringa oleifera seeds damages bacterial cell

membranes by causing membrane fusion. Langmuir, 31(15), 4496-4502.

Souza, P. F., Vasconcelos, I. M., Silva, F. D., Moreno, F. B., Monteiro-Moreira, A. C.,

Alencar, L. M., Abreu, A. S., Sousa, J. S. & Oliveira, J. T. (2016) A 2S albumin from

the seed cake of Ricinus communis inhibits trypsin and has strong antibacterial activity

against human pathogenic bacteria. Journal of Natural Products, 79(10), 2423-2431.

Sreerama, N., Venyaminov, S. Y. U. & Woody, R. W. (1999) Estimation of the number of

α‐helical and β‐strand segments in proteins using circular dichroism spectroscopy.

Protein Science, 8(2), 370-380.

Sreerama, N. & Woody, R. W. (2000) Estimation of protein secondary structure from

circular dichroism spectra: Comparison of CONTIN, SELCON, and CDSSTR methods

with an expanded reference set. Analytical Biochemistry, 287(2), 252-260.

Sreerama, N. & Woody, R. W. (2004) Computation and analysis of protein circular

dichroism spectra. Methods in Enzymology, 383, 318-345.

Strömstedt, A. A., Wessman, P., Ringstad, L., Edwards, K. & Malmsten, M. (2007) Effect

of lipid headgroup composition on the interaction between melittin and lipid bilayers.

Journal of Colloid and Interface Science, 311(1), 59-69.

Tai, S., Lee, T., Tsai, C., Yiu, T.-J. & Tzen, J. (2001) Expression pattern and deposition of

three storage proteins, 11S globulin, 2S albumin and 7S globulin in maturing sesame

seeds. Plant Physiology and Biochemistry, 39(11), 981-992.

Tormakangas, K., Hadlington, J. L., Pimpl, P., Hillmer, S., Brandizzi, F., Teeri, T. H. &

Denecke, J. (2001) A vacuolar sorting domain may also influence the way in which

proteins leave the endoplasmic reticulum. The Plant Cell, 13(9), 2021-2032.

120

Tseng, H.-K. & Perfect, J. R. (2011) Strategies to manage antifungal drug resistance.

Expert Opinion on Pharmacotherapy, 12(2), 241-241-256.

Vaccaro, A. M., Ciaffoni, F., Tatti, M., Salvioli, R., Barca, A., Tognozzi, D. & Scerch, C.

(1995a) pH-dependent conformational properties of saposins and their interactions with

phospholipid membranes. Journal of Biological Chemistry, 270(51), 30576-30580.

Vaccaro, A. M., Salvioli, R., Barca, A., Tatti, M., Ciaffoni, F., Maras, B., Siciliano, R.,

Zappacosta, F., Amoresano, A. & Pucci, P. (1995b) Structural analysis of saposin C and

B. Journal of Biological Chemistry, 270(17), 9953-9960.

Vaccaro, A. M., Tatti, M., Ciaffoni, F., Salvioli, R., Serafino, A. & Barca, A. (1994)

Saposin C induces pH-dependent destabilization and fusion of phosphatidylserine-

containing vesicles. FEBS Letters, 349(2), 181-186.

Vandenburg, Y. R., Smith, B. D., Biron, E. & Voyer, N. (2002) Membrane disruption

ability of facially amphiphilic helical peptides. Chemical Communications, 2002(16),

1694-1695.

Vivian, J. T. & Callis, P. R. (2001) Mechanisms of tryptophan fluorescence shifts in

proteins. Biophysical Journal, 80(5), 2093-2109.

Wang, Y., Grabowski, G. A. & Qi, X. (2003) Phospholipid vesicle fusion induced by

saposin C. Archives of Biochemistry and Biophysics, 415(1), 43-53.

Whitmore, L. & Wallace, B. A. (2008) Protein secondary structure analyses from circular

dichroism spectroscopy: methods and reference databases. Biopolymers, 89(5), 392-

400.

Williams, T. L., Day, I. J. & Serpell, L. C. (2010) The Effect of Alzheimer’s Aβ

Aggregation State on the Permeation of Biomimetic Lipid Vesicles. Langmuir, 26(22),

17260-17268.

Yang, Y., Heo, P., Kong, B., Park, J.-B., Jung, Y.-H., Shin, J., Jeong, C. & Kweon, D.-H.

(2015) Dynamic light scattering analysis of SNARE-driven membrane fusion and the

effects of SNARE-binding flavonoids. Biochemical and Biophysical Research

Communications, 465(4), 864-870.

Ye, Y. & Godzik, A. (2003) Flexible structure alignment by chaining aligned fragment

pairs allowing twists. Bioinformatics, 19(Suppl 2), ii246-ii255.

121

Yip, C. M. & McLaurin, J. (2001) Amyloid-β Peptide Assembly: A Critical Step in

Fibrillogenesis and Membrane Disruption. Biophysical Journal, 80(3), 1359-1371.

Zhai, Y. & Saier, M. H. (2000) The amoebapore superfamily. Biochimica et Biophysica

Acta - Reviews on Biomembranes, 1469(2), 87-99.

122

3.7 Appendices

3.7.A Considerations for using DLS to characterize vesicle fusion - As fusogenic

activity by PSI on PL vesicles proceeds, samples become increasingly polydispersed, both

in terms of multiple sizes appearing and disappearing within the vesicle population (not

only multiple sub-populations but also a relatively dramatic increase to the particle

diameter size range), and in terms of newly formed aggregates, thereby further increasing

the complexity of light scatter by the overall population (Pencer & Hallett, 2003).

Deviation from the initial vesicle population uniformity of size and particle shape

(spherical) eventually becomes too complex for correlation of light intensity and position

across the detector with the dynamics of a population of particles (Pencer & Hallett, 2003).

PSI-induced vesicle size change can be detectable within minutes, and therefore, the use

of lengthy measurement times and multiple incident light angles (necessary for accurate

size determination in polydispersed populations) are precluded. Therefore, when

considering DLS data for use in vesicle fusion assays, the data must be treated as suitable

for qualitative characterization of relative changes within a controlled experiment, but

should not be compared across platforms or experimental designs.

3.7.B Stability of large unilamellar vesicles - In terms of vesicle stability for stock LUV

preparations, Z-average diameter was monitored over long periods for test batches. Figure

3.16 outlines data from monitoring vesicle stability far past storage times used for

experiments, unexpectedly revealing relatively small fluctuations in average diameter at 38

days (1:1 POPE:POPS LUVs) and 50 days (1:1:1 POPC:POPE:POPS LUVs) post-

extrusion stored in the dark at ambient temperature. Negative control LUV samples post-

experiment were also randomly verified for size by DLS prior to being discarded (data not

123

Figure 3.16: Stability testing on LUV stocks after storage at ambient temperature

124

shown) to ensure that assay conditions did not cause size distortions (e.g., shear stress from

shaking polyethylene microplate-based leakage assays, pipetting in and out of cuvettes and

stock tubes, repeated UV light exposure during time trial assays). These tests were intended

as assurance of vesicle stability; that fusion assays and cryo-TEM experiments were in the

context of substrate that is not prone to destabilizing without PSI or other intervention.

Also anecdotally, but potentially important to many researchers, it has been noted

repeatedly that LUV preparations used in the present study are more stable when

refrigeration temperatures are avoided and the authors recommend re-evaluating protocols

that call for refrigeration during LUV storage.

3.7.C Considerations for comparing leakage rates at low pH values - Although the

results in Figure 3.4 are shown normalized across groups (to the highest rate measured),

we caution that comparison of calcein fluorescence rates of change between different pH

conditions is troublesome due to low fluorescence efficiencies at low pH. The signal ranges

from negative control (intact LUVs) to positive control (full calcein release) are essentially

equal and highest at pH 7.4 and pH 6.2 whereas range is approximately 4-fold lower at pH

4.5 and 12-fold lower at pH 3.0. These values were variable, but of typical LUV

preparations and assays over many experiments. Noise introduced greater relative error to

pH 3.0 rates compared to those at pH 4.5. Furthermore, fluorescence emission as a function

of concentration was not consistent between pH values below 6.2 presumably reflecting

the effects of varying quantum yield on fluorescence emission itself as well as self-

quenching efficiency / inner filter effect. At constant pH, the latter can be ignored for

measuring initial leakage rates due to the leaked calcein concentration in the reaction

volume being low (Kendall & MacDonald, 1982), however, comparison of leakage rates

125

for reactions with self-quenching fluorophore reporter at different quantum yields (i.e., pH

3.0 vs, pH 4.5) is likely unreliable quantitatively, particularly at pH 3.0. The important

finding in the context of the present study is that PSI appears to have strong PL bilayer

perturbation activity at both pH 3.0 and pH 4.5.

126

Chapter 4: Comparative Structure-Function Characterization

of the Saposin-Like Domains from Potato, Barley, Cardoon

Thistle and Arabidopsis Aspartic Proteases

Note: The findings and content of following chapter were submitted (January 2017) as a

revised manuscript to Biochimica et Biophysica Acta - Biomembranes for review.

4.1 Abstract

The present study characterized the aspartic protease saposin-like domains of four plant

species in terms of bilayer fusion, bilayer perturbation, and structure pH-dependence.

Recombinant saposin-like domains from Solanum tuberosum (potato), Hordeum vulgare

L. (barley), Cynara cardunculus (Cardoon thistle) and Arabidopsis thaliana were

compared revealing that all had vesicle leakage activities against simple phospholipid

bilayer vesicles with the relative rates as Arabidopsis > barley > Cardoon > potato. When

compared against a bilayer composed of a vacuole-like phospholipid mixture, leakage was

more than five times higher for potato saposin compared to the others. In terms of fusogenic

activity, all showed faster and larger particle size increases relative to potato saposin,

particularly for barley and Arabidopsis saposins. Bilayer fusion assays in reducing

conditions caused clear alterations in the fusion product profiles except for the case of

Arabidopsis saposin which was virtually unchanged. Secondary structure contents were

similar across all four proteins under different pH conditions, although Cardoon saposin

appeared to have higher overall helix structure. In terms of Trp emission for saposin

interactions with bilayer, rates for reaching equilibrium indicated that more than half of the

structural rearrangement for saposin-like domain bilayer interactions occurred in less than

two minutes upon encountering vesicles under experimental conditions. Overall, the

127

present findings serve as a foundation for future studies seeking to attribute functionality

differences to bilayer interacting structural motifs or other variants among the plant saposin

domains as well as between the domains and other saposin-like proteins.

4.2 Introduction

For several decades, the relationship in food plants post-harvest (Yomo & Srinivasan,

1973) as well as during seed maturation (Jacobsen & Varner, 1967) and germination (Garg

& Virupaksha, 1970) with increased proteolytic activity has been noted. Also of importance

is the association of acid protease activity within the lysosome (Yatsu & Jacks, 1968), one

of two functions of a type of plant acid protease-associated protein that is the focus of the

present study. Protein degradation and processing being an important function of this

organelle, more than two dozen proteases are in fact found in the plant vacuole

(Arabidopsis thaliana) (Carter et al, 2004). In addition to cysteine proteases (McGrath,

1999) and serine proteases (Breddam, 1986), aspartic proteases (AP) are part of the

ensemble of vacuolar enzymes, analogous to its lysosomal equivalent in animal cells

(Marty, 1999).

The two earliest structures to be reported for plant APs were phytepsin (Kervinen et al,

1999) from barley (Hordeum vulgare L.) and cardosin A (Frazão et al, 1999) from the

flowers of Cardoon (Cynara cardunculus L.), the latter being a source of vegetable rennet

specifically due to cardosin milk-clotting activity (Veríssimo et al, 1995). Phytepsin,

crystalized in its zymogen form, displayed the usual bilobal AP tertiary structure, and also

revealed the first structural detail of a domain uniquely found in plant APs termed the plant-

specific insert (PSI) or plant-specific sequence (PSS) (Runeberg‐Roos et al, 1991;

128

Sarkkinen et al, 1992), most closely related to saposins in terms of primary, and predicted

secondary and tertiary structures (Guruprasad et al, 1994). This predicted structural

similarity was first verified for phytepsin (Kervinen et al, 1999) and more recently for the

PSI of Solanum tuberosum AP (Bryksa et al, 2011) for a recombinantly expressed product

sans its usual AP structural partner. The particular interest in the latter stems from its well-

demonstrated anti-pathogen activities both in vivo (Guevara et al, 2002; Mendieta et al,

2006; Pagano et al, 2007) and in vitro (Muñoz et al, 2010).

The structure-function relationships of the saposin-like domains of plant aspartic

proteases have only recently begun to be unraveled despite the large body of research on

other saposin-like proteins, as well as their importance to agricultural success and stability,

and food security by way of their roles in plant resistance to pathogens. Compared to many

saposin-like proteins (SAPLIP) (Bruhn, 2005; Vaccaro et al, 1999), particularly NK-lysin

and surfactant protein B and others having long standing recognition for their critical roles

in health and potential directed usage in disease treatments (Gaspar et al, 2013), the

saposin-like domains of plant aspartic proteases remain largely untouched with respect to

their biochemical properties. Throughout the saposin literature, a wide array of different

PLs, vesicle sizes and temperatures can be readily found. It was reported that PL

composition has an important impact on bilayer disruption (Muñoz et al, 2011), rendering

it difficult to compare across platforms. Lastly, temperature is potentially important to PL

physical state so variability between projects could be problematic even if studying the

same PL vesicles.

Generally, delineating structure-function relationships within a given group of proteins

is aided by the existence of highly similar proteins having small sequence differences,

129

providing an experimental framework in which to operate, and such is the situation

presented by plant AP PSIs. The present study compared the PSIs of four species, namely

Solanum tuberosum, Hordeum vulgare L., Cynara cardunculus L. and Arabidopsis

thaliana, in terms of their bilayer interactions and protein structural features. The

investigation appears to be the first comparative study of multiple plant saposins wherein

all received the same treatments, environmental conditions and LUVs. The results highlight

some basic functional differences among the four PSIs despite their highly similar

structures.

4.3 Experimental Procedures

4.3.1 Materials - Synthetic genes for each PSI were optimized for expression in E. coli

and purchased from Mr. Gene GmbH (Regensburg, Germany). Plasmid pET32b(+), E. coli

Rosetta-gami B (DE3)pLysS, and u-MACTM columns were obtained from EMD

Biosciences (San Diego, CA, USA). E. coli TOP10F’ was from Invitrogen (San Diego,

CA, USA). GenEluteTM Plasmid Miniprep Kit was obtained from Sigma-Aldrich Co. (St.

Louis, MO, USA). The QIAquick® PCR Purification Kit and QIAquick® Gel Extraction

Kit were from Qiagen (Germantown, MD, USA). Restriction enzymes, T4 DNA ligase and

Pfu DNA polymerase were obtained from Fermentas Life Sciences (Burlington, ON,

Canada). Primers were synthesized by Sigma Genosys (Oakville, ON, Canada) and

thrombin was purchased from Fisher Scientific Co. (Ottawa, ON, Canada). The RPC

column was from GE Healthcare (Piscataway, NJ, USA). Phospholipids (PL) were from

Avanti Polar Lipids (Alabaster, AL, USA).

130

4.3.2 PSI Expression and Purification - Sub-cloning the Solanum tuberosum AP gene

into pET23b was described previously as was the method used for recombinant expression

and purification for all PSIs in the present study (Bryksa et al, 2011). Primers used for sub-

cloning the three other PSI synthetic genes for the present study were:

Phytepsin PSI: 5’-atccatggcggtggttagtcaggagtgtaaaacg, and

5’-atctcgagttacggcagacggttacacagtt;

Cardosin A PSI: 5’-atccatggcggtgatgaaccagcagtgtaaaac, and

5’–atctcgagttagctcaggtgttcacacagct

AtAP PSI: 5’-atccatggcggtggtttctcagcagtgtaaaacg, and

5’-atctcgagttaaggcagacgctcacacagt

4.3.3 Preparation of Large Unilamellar Vesicles (LUV) - LUV stocks were prepared

as per (Bryksa et al, 2011) with the exception that 80 mM calcein / 140 mM NaCl / 25 mM

Na-acetate pH 4.5 was used in order to suspend dried PL mixtures at 42º C after removal

of storage solvent under nitrogen flush. Phospholipids used were 1-palmitoyl-2-oleoyl-sn-

glycero-3-phosphocholine (POPC), 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoethanol-

amine (POPE) and 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoserine (POPS) for 1:1

POPE:POPS or 1:1:1 POPC:POPE:POPS. For vacuole-like PL mixture vesicles, the

relative amounts of each PL are detailed in Table 4.1.

4.3.4 Circular Dichroism Spectropolarimetry (CD) - CD analysis of PSI secondary

structure was carried out using a Jasco J-810 spectropolarimeter (Jasco Inc., Easton, MD,

USA). PSIs (10 µM) were scanned over 196-260 nm at 100 nm/min, 0.5 s response,

standard sensitivity, and at ambient temperature using a 1 mm pathlength quartz cell.

4.3.5 LUV Disruption Assays - PSI-caused perturbation of LUVs was measured by

calcein-loaded vesicles and leakage was detected using a Victor2 1420 Multilabel Counter

(Perkin Elmer, Waltham, MA, USA) or a Shimadzu RF-540 spectrofluorophotometer

131

Table 4.1: Phospholipid composition used for preparing LUVmix. The composition

mimics those of barley root endoplasmic reticulum, tonoplast and Golgi membrane bilayers (Brown

& DuPont, 1989). The commercially available PLs used in the present study were selected for their

stabilities at ambient temperature and shelf-life / cost, in addition to their similarities to the lipid

mixtures of barley root PLs.

132

(Shimadzu Corporation, Kyoto, Japan) at 25° C. 200 µL reactions were set up in 96-well

microplates or a quartz ultra-microcuvette with varying concentrations of LUVs, 500 nM

PSI and 140 mM NaCl / 25 mM Na-acetate pH 4.5. Leakage was detected using excitation

at 385 nm and emission at 435 nm. Endpoints were measured by incubating LUVs in 0.5%

triton for each condition. Non-linear regression analyses were done using GraphPad Prism

4 (GraphPad Software Inc., La Jolla, CA, USA).

4.3.6 Bilayer Fusion Assays - Ten μM PSI or peptide was incubated with 100 nm LUVs

(100 μM total PL; 1:1 molar ratio of POPE:POPS) at 25° C in either 25 mM Na-acetate /

140 mM NaCl pH 4.5 or 5 mM Na-phosphate / 140 mM NaCl pH 7.4 (control). Mixtures

were then monitored for changes in average LUV size by dynamic light scattering in a

Malvern Zetasizer Nano-S (Malvern Instruments, Malvern, Worcestershire, UK) using a

disposable polystyrene 1.5 mL semi-microcuvette. Three consecutive measurements of

five 30 s runs each were averaged using the refractive index for polystyrene.

4.3.7 Tryptophan Intrinsic Fluorescence Spectrometry - Fluorescence spectra were

recorded using a Shimadzu RF-540 spectrofluorophotometer (Shimadzu Corporation,

Kyoto, Japan) with a 1-cm quartz three-sided ultra-micro cuvette in a water-circulating

temperature-controlled cell holder at 25° C.  The settings used were λexcitation 295 nm with

3 nm slit width and λemission scan 300-400 nm with 3 nm or 5 nm slit width. PSI monomer-

dimer characterization was done at 10 μM, and PSI/peptide-bilayer scans used 8.5 μM

protein and 100 nm LUVs (100 µM total PL; 1:1:1 molar ratio of POPC:POPE:POPS).

133

4.4 Results and Discussion

4.4.1 Primary Structure Comparison - Primary structure identity as well as charge

inversions and regional differences in hydrophobicity were assessed by multiple sequence

alignments of StAP, phytepsin, cardosin A and AtAP (Figure 4.1.A) using PRALINE

(Simossis & Heringa, 2005). Among the four sequences, identity was 0.61 with a clear

majority of the sequences indicating highly conserved regions in terms of both residue type

and hydrophobicity. An 11-residue near perfectly conserved sequence between the four

PSI species (indicated with a red box in Figure 4.1.B) is located centrally within the C-

terminal half of the PSIs, surrounding the singular requisite Trp77 (StAP PSI numbering).

Among the four PSIs, some noteworthy differences indicated in Figure 4.1.B are the N-

terminal domain Trp unique to StAP (green diamond) as well as various unmatched or

inverted charges (red or orange diamonds, respectively) for acidic and basic residues.

The sequences of the canonical SAPLIPs saposin B and saposin C were aligned with

the swapped N- and C-terminal halves of StAP PSI (Figure 4.1.C). Relative to other

SAPLIPs, helices 1 and 2 of the N-terminal half of saposin are structurally equivalent to

PSI helices 3 and 4 of its C-terminal half while H3 and H4 of saposin are equivalent to PSI

H1 and H2 (Figure 4.1.C), hence the term “swaposin” (Heinemann & Hahn, 1995). The

alignment produced high scoring sequence conservation (49% sequence identity), and

secondary structure alignment. Since the vast majority of research to date on SAPLIPs has

been done on non-plant sources, most available information must be applied to PSIs

through saposin-swaposin residue equivalencies indicated in Figure 4.1.C.

134

Figure 4.1: PSI primary structure. (A) PRALINE multiple sequence alignments indicating

amino acid conservation (top) and type/hydrophobicity (bottom) for AtAP (1), phytepsin (2),

cardosin A (3) and StAP (4) PSIs. (B) Multiple sequence alignment generated using CLUSTRAL

W for AtAP (1), phytepsin (2), cardosin A (3) and StAP (4) PSIs, identifying charge differences

(red diamonds), charge inversions (orange diamonds) and the additional Trp exclusive to StAP PSI

(green diamond). (C) PRALINE multiple sequence alignment for human saposin B (1), human

saposin C (2) and StAP PSI (3) coloured based on sequence conservation. StAP PSI is presented

with its N- and C-terminal halves swapped to align with the saposins.

135

4.4.2 Bilayer Disruption and Fusion - StAP, phytepsin, cardosin A and AtAP PSIs

were initially compared based upon calcein leakage rates for POPE:POPS LUVs at three

PL concentrations in buffered saline pH 4.5 (Figure 4.2.A). No statistically significant

differences in maximum leakage rates were measured between PSI species at 100 µM and

20 µM PL, respectively (P>0.05), whereas all pairs were significantly different (P≤0.05),

respectively, at 500 µM PL. Although all were within an order of magnitude, StAP PSI had

the lowest membrane perturbation activity, up to 3-fold lower relative to the other 3 species.

The present study principally used vesicles composed of simple PL mixtures (1:1

POPE:POPS and 1:1:1 POPC:POPE:POPS) as previously (Bryksa et al, 2011; Bryksa &

Yada, 2017), however, being a comparison between multiple PSIs, a new PL blend that

mimics the PL composition of barley vacuolar plasma membrane (Brown & DuPont,

1989), as outlined above in Table 4.1, was incorporated for making LUVs. As a set of

parallel measurements to the aforementioned simpler PL blends, the vacuole-like bilayer

was meant to provide a balance between a more realistic substrate for PSIs and simplicity

(i.e., no membrane proteins, plant cholesterol, carbohydrates) for optimal signal-to-noise

and repeatability.

Although slower than the rates for POPE:POPS LUVs, PSIs were each active against

the vacuole-like LUVmix bilayer vesicles, with StAP having significantly higher leakage

rate (P≤0.05, respectively) than the other three PSI species (Figure 4.2.B). Thus, the

differences in leakage activity among the PSIs were somewhat unclear. As noted above for

POPE:POPS vesicles, the PSIs had similar activity levels at 20 µM and 100 µM PL, but

significantly higher (P≤0.05) leakage for cardosin A, phytepsin and AtAP PSIs, in that

order, at 500 µM PL. Contrasting this were the leakage results for LUVmix where StAP PSI

136

Figure 4.2: PSI-induced PL bilayer vesicle leakage in isoionic saline buffers at 25° C.

Rates were calculated and expressed as relative to the highest data point measured across all data

sets. (A) StAP, phytepsin, cardosin A and AtAP PSIs compared at three POPE:POPS

concentrations. (B) The same compared using the vacuole-like PL composition LUVmix at 100 µM

PL. Error bars indicate +/- one standard deviation.

137

had much higher activity (Figure 4.2) with the other three PSIs having less than 15% of the

relative rate of leakage.

PSI-induced bilayer fusion was characterized by monitoring POPE:POPS LUV size at

25° C in buffered saline pH 4.5 with and without 10 mM DTT (Figures 4.3 and 4.4,

respectively). Assays were ended when vesicle size populations became overly

polydisperse for DLS measurements. Under native conditions, phytepsin, cardosin and

AtAP PSIs all caused markedly faster onset of larger and more complete transitions of the

vesicle size populations (Figure 4.3). Within 30 min, both phytepsin and AtAP PSIs caused

100% of the original size population to convert to larger apparent sizes. By contrast, vesicle

size increased more gradually for 95-98% of the overall population upon incubation with

StAP PSI (i.e., the initial peak broadened incrementally), and 2-5% of the overall

population formed relatively large structures (Figure 4.3). Although overall cardosin A

PSI-induced size shifting was slower compared to phytepsin and AtAP, it nonetheless also

caused enlargement of the entire original population by 20 h. Notably, relatively narrow

vesicle size populations were maintained throughout the apparent fusion process (Figure

4.4.C). Also, the action of both cardosin A and phytepsin PSIs on POPE:POPS LUVs

resulted in a lack of fused structures much greater in size than 1000 nm. In reducing

conditions, fusogenic activity appeared to be virtually unaffected for phytepsin and

cardosin A PSIs (Figure 4.3.B-C compared to Figure 4.4.B-C) whereas AtAP activity

resulted in a wider range of vesicle sizes relative to non-reducing conditions. StAP PSI

action was most affected by the absence of its disulfide bonds compared to the other PSIs.

Although StAP PSI produced the narrowest overall vesicle size range in native conditions

(Figure 4.3), StAP, phytepsin and AtAP PSIs all produced similar fusion profiles upon

138

Figure 4.3: Bilayer fusion of 100 µM POPE:POPS as 100 nm LUVs. Size was monitored

by DLS 25° C upon incubation with 10 µM of either StAP, phytepsin, cardosin A or AtAP PSI.

Three consecutive measurements of five 30 s runs each were averaged using the refractive index

for polystyrene.

139

Figure 4.4: Bilayer fusion of 100 µM POPE:POPS as 100 nm LUVs in reducing

conditions. Size was monitored by DLS 25° C upon incubation with 10 µM of either StAP,

phytepsin, cardosin A or AtAP PS. Three consecutive measurements of five 30 s runs each were

averaged using the refractive index for polystyrene.

140

elimination of their disulfide bonds (Figure 4.4). This may implicate a common tertiary

structural feature dependent on the PSI disulfide bonds that, when lost, results in a loss of

differentiation of function/specificity.

Fusion assays were additionally carried out for vacuole-like LUVmix vesicles (Figure

4.5), where all four PSIs were indicated to have fusogenic activity, and phytepsin PSI being

the fastest. Due to the slower fusion rates relative to POPE:POPS LUVs, Figure 5 utilizes

a linear diameter scale to better visualize changes, and time points at 5, 15 and 30 min are

not shown to improve clarity in these figures. The overall lower bilayer effects compared

to the simpler PL mixes was in general agreement with (Muñoz et al, 2011). The lower

particle sizes obtained were not surprising in that these vesicles contain approximately half

of the total negative charge compared to 1:1 PE:PS bilayers, and therefore, lowered contact

incidence would be expected. Phytepsin displayed the highest fusogenic activity, possibly

reflective of the fact that the LUVmix PL content is based upon barley root vacuolar plasma

membranes (Brown & DuPont, 1989). After 9 h incubation, fusion mixtures were reduced

with DTT and left for 24 h to verify that PSI-induced vesicle size increases for LUVmix

vesicles remained stable (i.e., no unexpected profile changes) despite reduction of PSI

disulfide bonds. This verification was done in the context of the findings that reduction of

PSI disulfide bonds did not prevent vesicle fusion (Fig. 4.4 and 4.5). Similarly, vesicle size

increase did not appear to have deviated, qualitatively, from the trend observed through 9

h in non-reducing conditions for the four PSIs.

Since the vacuole-like LUVmix vesicles had not been studied previously, the stability of

vesicle stocks were assessed upon a 3-month storage period at ambient temperature in

141

Figure 4.5: Bilayer fusion of 100 µM LUVmix as 100 nm LUVs. Size was monitored by

DLS 25° C upon incubation with 10 µM of either StAP, phytepsin, cardosin A or AtAP PSI.

Three consecutive measurements of five 30 s runs each were averaged using the refractive index

for polystyrene.

142

buffered saline pH 4.5. No detectable size changes were noted for two separately prepared

batches (i.e., 126 vs. 127 nm), while the vesicles leaked only ~16%, indicating that LUVmix

vesicles are highly stable. Although more susceptible to action by PSIs, 1:1 POPE:POPS

and POPC:POPE:POPS vesicles were also found to be stable, to a lesser extent, over five

weeks (i.e., size increases of ~3% and ~7%, respectively), reinforcing that effects on

bilayers observed within experimental timeframes (minutes or hours) were strictly due to

PSI.

4.4.3 Comparison of pH-dependence of secondary structure - StAP, phytepsin,

cardosin A and AtAP PSIs were compared qualitatively in terms of secondary structure in

buffered saline pH 3.0, 4.5, 6.2, and 7.4 (iso-ionic across the four buffers). The CD spectral

changes for the respective PSIs as well as inter-species comparisons are presented in

Appendices A-B. For all four PSIs, helix content increased with decreasing pH, evidenced

by the stronger ellipticity measurements at 222 and 208 nm. In terms of secondary structure

differences between the PSIs across pH conditions, cardosin A PSI appeared to have

consistently higher helix content while phytepsin appeared to have consistently lower

overall helix content across all pH values, based on the relative magnitudes of the negative

peaks at 222 and 208 nm. With the possible exception of phytepsin, the distinct secondary

structure states between neutral and acidic pH, previously reported for StAP PSI (Bryksa

et al, 2011), were consistent with cardosin A and AtAP PSIs (see Appendices A-B) of the

present study. Despite the presence of an extra Trp in the N-terminal portion of StAP PSI,

which is suspected to influence secondary structure (Bryksa et al, 2011; Lai & Tamm,

2007), CD scans and patterns of CD spectral changes over the pH range studied did not

produce clear differences. Although Trp is important to the fusogenic helical boomerang

143

configuration (Lai & Tamm, 2007), the overall helical content would not necessarily be

appreciably altered, possibly explaining the apparent lack of CD spectral characteristics

between StAP and the other PSIs.

4.4.4 Intrinsic Trp fluorescence in solution and PSI-bilayer interactions - Intrinsic

Trp fluorescence was used to detect evidence for monomer-dimer equilibrium as reported

for StAP PSI (Chapter 3). StAP PSI was to thus serve as a known benchmark since it had

been measured by analytical centrifugation (Chapter 3). The emission scans are shown in

Figure 6 where the distinct two-state pH-dependent curve magnitudes for StAP PSI

coincide with its inactive/active pH profile as well as confirmation by analytical

centrifugation (Bryksa et al, 2011). The optimal wavelength for emission among the spectra

were also plotted for detecting red- or blue-shift as an indicator of changes in Trp

environment, showing that λmax was unchanged across all datasets (P>0.05; data not

shown). Although the emission spectra in Figure 4.6 were all measured under the same

conditions and for the same PSI concentration (8.5 µM), the absolute fluorescence emission

for StAP at its lowest point (pH 7.4) was still almost double the highest signal strength

emitted at any pH for the other three PSIs due to the presence of a second Trp. The emission

differential was in spite of using an emission slit width of 5 nm for phytepsin, cardosin and

AtAP PSIs instead of 3 nm used for StAP (Figure 4.6) to improve signal-to-noise as

emission peaked approximately 2.5-fold lower at 3 nm slit width compared to that at 5 nm.

The relatively low emission signals for phytepsin, cardosin A and AtAP PSIs were

consistent with saposin C (Wang et al, 2003), but problematic for trying to characterize an

increase in signal that was small (~10% increase for the dimer relative to the monomer) for

StAP PSI (Figure 4.6). Overall, among the respective emission spectra, no clear evidence

144

Figure 4.6: Intrinsic Trp fluorescence emission spectra for StAP, phytepsin, cardosin

A and AtAP PSIs in buffered saline at varying pH. The respective λmax measurements at

the indicated pH values for each PSI are inset, respectively. Error bars indicate +/- standard error.

145

was presented for an StAP-like monomer-dimer arrangement among the other three PSIs

that followed an interpretable pattern, however, such an arrangement is not precluded

either. There were differences within the data sets shown in Figure 4.6 (i.e., higher emission

at pH 7.4 for cardosin A PSI; lower emission at pH 4.5 for phytepsin PSI; and higher

emission for pH 3.0 AtAP PSI). Among these results, the emission increase at lower pH

for AtAP was the only one that fit the monomer-dimer model given by StAP PSI, however,

this could not be interpreted in the context of phytepsin and cardosin A PSIs having

seemingly erroneous emission spectra at pH 4.5 and pH 7.4, respectively (i.e., phytepsin

monomer at pH 4.5 and cardosin A dimer at pH 7.4). If these emission changes are

reflections of Trp environment then they could not be explained. A possible confounding

factor may be differences in local structure proximate to Trp77 (the Trp common to all 4

PSIs) such that solvent exposure differences not directly related to monomer-dimer status

cause altered fluorescence emission. Further study of the monomer-dimer status of the

phytepsin, cardosin A and AtAP PSIs in solution will require a different experimental

design. At present, structural studies are ongoing which will aid in understanding local

effects of Trp77.

Spectral changes in Trp fluorescence emission upon encountering 1:1:1

POPC:POPE:POPS LUV bilayer were also compared (Figure 4.7). No λmax shifts were

observed for the four PSIs. Kinetic analyses of Trp fluorescence increase were done as a

measure of the rate at which PSI equilibrated such that spectra were collected over time

periods that approached the endpoints for increasing emission (preliminary spectra were

collected well past detectable changes in emission to define these amounts of time; 300–

650 s). Data were then normalized to their respective endpoints so that signal strength

146

Figure 4.7: Trp fluorescence emission of PSIs upon incubation with anionic bilayer

vesicles. (A) Intrinsic Trp fluorescence emission spectra for 10 µM PSI with 100 µM 1:1:1

POPC:POPE:POPS LUVs. (B) Kinetics of emission increase indicating equilibration of

PSI with bilayer (one-phase non-linear association).

147

differences between samples were comparable. Individual plots are shown in Figure 4.7.A,

and the kinetics for increasing maximum emission measurements for the four PSIs are

summarized in Figure 4.7.B. After several scans of consistently increasing emission over

the first 200–500 s, spectra began to overlap and eventually superposed, indicating that a

new equilibrium state had been reached. The equilibration processes for the four PSIs were

calculated for the increase in maximum emission over time yielding the rates listed in the

table inset within Figure 4.7.B, suggesting that equilibration was 50% complete within the

first two minutes for all PSIs. StAP PSI half-time was 52+/-3 s, and approximately 90% of

the change occurred within the first 3 min. In a previous study (Chapter 3) which focused

on StAP PSI, timed CD scans in the presence of vesicles was also complete around the 4.5

min mark. The possible link between these two data (i.e., secondary structure change on a

similar timeline as tertiary/quaternary structure environment change) will require further

investigation.

The lack of λmax shifting observed for the four PSIs, a result that was previously found

for StAP PSI alone (Chapter 3), was unexpected considering the presumed insertion of Trp

into the bilayer (Popovic et al, 2012; Qi & Grabowski, 2001; Wang et al, 2003; Willis et

al, 2011). Furthermore, all measured λmax values were notably low (blue-shifted; elevated

Trp environment hydrophobicity) compared to previous results for saposins (Popovic et al,

2012; Qi & Grabowski, 2001; Wang et al, 2003). The present findings provide new context

to the emission spectral anomalies previously reported (Chapter 3) where it was suggested

that a suspected Internal Stark Effect (Vivian & Callis, 2001) was manifesting within the

H3 region of StAP PSI based upon: (i) the lack of blue shifting in emission spectra despite

the presence of a membrane penetrating motif in a proven bilayer-active protein despite

148

exposure to a favorable bilayer target; and (ii) the presence of Glu19 neighboring Trp18.

Since StAP PSI contains two Trp, it was postulated that the Glu-Trp-induced red shift and

a concomitant bilayer-Trp-induced blue shift, thereby offsetting expected change in λmax.

Although the above may be correct for StAP PSI, the present findings from a comparative

study that included three single-Trp PSIs may further clarify the phenomena in that

phytepsin, cardosin A and AtAP PSIs all contain Trp77 as their only source of Trp

fluorescence emission (via excitation at 295 nm) and none have a neighbouring charged

residue. Therefore, all of the elements that made the Internal Stark Effect a possible

explanation for the lack of blue shift in emission spectra of StAP PSI are absent for

phytepsin, cardosin A and AtAP PSIs, yet each showed the same lack of blue shift.

Speculatively, an alternative explanation for this apparent unchanging high-hydrophobicity

status may be that Trp77, upon encountering bilayer, essentially transitions from a dimer

hydrophobic pocket (Bryksa et al, 2011) to bilayer interior, effectively trading one highly

hydrophobic environment for another. The PSI Trp77-bilayer relationship will be

characterized by ongoing protein-bilayer investigations at the atomic scale. Although the

study of the saposin-like domains of plant APs is in its infancy, the subtle primary structure

distinctions and accompanying subtle differences in bilayer functionalities between the

PSIs of the present study are suggestive of the potential for exploiting swaposins to gain

detailed insight into membrane-active protein sequences, essential for emerging SAPLIP-

related targeted delivery, cell penetrating and surface active biotechnologies (Fonseca et

al, 2009; Frauenfeld et al, 2016; Walther et al, 2016).

149

4.5 References

Breddam, K. (1986) Serine carboxypeptidases. A review. Carlsberg Research

Communications, 51(2), 83-128.

Brown, D. J. & DuPont, F. M. (1989) Lipid composition of plasma membranes and

endomembranes prepared from roots of barley (Hordeum vulgare L.) effects of salt.

Plant Physiology, 90(3), 955-961.

Bruhn, H. (2005) A short guided tour through functional and structural features of saposin-

like proteins. Biochemical Journal, 389, 249-257.

Bryksa, B. C., Bhaumik, P., Magracheva, E., DeMoura, D. C., Kurylowicz, M., Zdanov,

A., Dutcher, J. R., Wlodawer, A. & Yada, R. Y. (2011) Structure and mechanism of

the saposin-like domain of a plant aspartic protease. Journal of Biological Chemistry,

286(32), 28265-28275.

Carter, C., Pan, S., Zouhar, J., Avila, E. L., Girke, T. & Raikhel, N. V. (2004) The

vegetative vacuole proteome of Arabidopsis thaliana reveals predicted and unexpected

proteins. The Plant Cell, 16(12), 3285-3303.

Fonseca, S. B., Pereira, M. P. & Kelley, S. O. (2009) Recent advances in the use of cell-

penetrating peptides for medical and biological applications. Advanced Drug Delivery

Reviews, 61(11), 953-964.

Frauenfeld, J., Loving, R., Armache, J.-P., Sonnen, A. F. P., Guettou, F., Moberg, P., Zhu,

L., Jegerschold, C., Flayhan, A., Briggs, J. A. G., Garoff, H., Low, C., Cheng, Y. &

Nordlund, P. (2016) A saposin-lipoprotein nanoparticle system for membrane

proteins. Nature Methods, 13(4), 345-351.

Frazão, C., Bento, I., Costa, J., Soares, C. M., Veríssimo, P., Faro, C., Pires, E., Cooper, J.

& Carrondo, M. A. (1999) Crystal structure of cardosin A, a glycosylated and Arg-

Gly-Asp-containing aspartic proteinase from the flowers of Cynara cardunculus L.

Journal of Biological Chemistry, 274(39), 27694-27701.

Garg, G. K. & Virupaksha, T. K. (1970) Acid protease from germinated sorghum 2.

Substrate specificity with synthetic peptides and ribonuclease A. European Journal of

Biochemistry, 17(1), 13-18.

150

Gaspar, D., Veiga, A. S. & Castanho, M. A. R. B. (2013) From antimicrobial to anticancer

peptides. A review. New edge of antibiotic development: antimicrobial peptides and

corresponding resistance, 4, 294.

Guevara, M. G., Oliva, C. R., Huarte, M. & Daleo, G. R. (2002) An aspartic protease with

antimicrobial activity is induced after infection and wounding in intercellular fluids of

potato tubers. European Journal of Plant Pathology, 108(2), 131-137.

Guruprasad, K., Törmäkangas, K., Kervinen, J. & Blundell, T. L. (1994) Comparative

modelling of barley-grain aspartic proteinase: A structural rationale for observed

hydrolytic specificity. FEBS Letters, 352(2), 131-136.

Heinemann, U. & Hahn, M. (1995) Circular permutations of protein sequence: Not so rare?

Trends in Biochemical Sciences, 20(9), 349-350.

Jacobsen, J. V. & Varner, J. E. (1967) Gibberellic acid-induced synthesis of protease by

isolated aleurone layers of barley. Plant Physiology, 42(11), 1596-1600.

Kervinen, J., Tobin, G. J., Costa, J., Waugh, D. S., Wlodawer, A. & Zdanov, A. (1999)

Crystal structure of plant aspartic proteinase prophytepsin: inactivation and vacuolar

targeting. EMBO Journal, 18(14), 3947-3955.

Lai, A. L. & Tamm, L. K. (2007) Locking the kink in the influenza hemagglutinin fusion

domain structure. Journal of Biological Chemistry, 282(33), 23946-23956.

Marty, F. (1999) Plant Vacuoles. The Plant Cell, 11(4), 587-599.

McGrath, M. E. (1999) The lysosomal cysteine proteases. Annual Review of Biophysics

and Biomolecular Structure, 28(1), 181-204.

Mendieta, J. R., Pagano, M. R., Munoz, F. F., Daleo, G. R. & Guevara, M. G. (2006)

Antimicrobial activity of potato aspartic proteases (StAPs) involves membrane

permeabilization. Microbiology, 152, 2039-2047.

Muñoz, F., Palomares-Jerez, M. F., Daleo, G., Villalaín, J. & Guevara, M. G. (2011)

Cholesterol and membrane phospholipid compositions modulate the leakage capacity

of the swaposin domain from a potato aspartic protease (StAsp-PSI). Biochimica et

Biophysica Acta - Molecular and Cell Biology of Lipids, 1811(12), 1038-1044.

Muñoz, F. F., Mendieta, J. R., Pagano, M. R., Paggi, R. A., Daleo, G. R. & Guevara, M.

G. (2010) The swaposin-like domain of potato aspartic protease (StAsp-PSI) exerts

antimicrobial activity on plant and human pathogens. Peptides, 31(5), 777-785.

151

Pagano, M. R., Mendieta, J. R., Muñoz, F. F., Daleo, G. R. & Guevara, M. G. (2007) Roles

of glycosylation on the antifungal activity and apoplast accumulation of StAPs

(Solanum tuberosum aspartic proteases). International Journal of Biological

Macromolecules, 41(5), 512-520.

Popovic, K., Holyoake, J., Pomès, R. & Privé, G. G. (2012) Structure of saposin A

lipoprotein discs. Proceedings of the National Academy of Sciences, 109(8), 2908-

2912.

Qi, X. & Grabowski, G. A. (2001) Differential membrane interactions of saposins A and

C. Journal of Biological Chemistry, 276(29), 27010-27017.

Runeberg‐Roos, P., Tormakangas, K. & Östman, A. (1991) Primary structure of a barley‐

grain aspartic proteinase. European Journal of Biochemistry, 202(3), 1021-1027.

Sarkkinen, P., Kalkkinen, N., Tilgmann, C., Siuro, J., Kervinen, J. & Mikola, L. (1992)

Aspartic proteinase from barley grains is related to mammalian lysosomal cathepsin

D. Planta, 186(3), 317-323.

Simossis, V. A. & Heringa, J. (2005) PRALINE: a multiple sequence alignment toolbox

that integrates homology-extended and secondary structure information. Nucleic Acids

Research, 33(Supplement 2), W289-W294.

Vaccaro, A. M., Salvioli, R., Tatti, M. & Ciaffoni, F. (1999) Saposins and their interaction

with lipids. Neurochemical Research, 24(2), 307-314.

Veríssimo, P., Esteves, C., Faro, C. & Pires, E. (1995) The vegetable rennet of Cynara

cardunculus L. contains two proteinases with chymosin and pepsin-like specificities.

Biotechnology Letters, 17(6), 621-626.

Vivian, J. T. & Callis, P. R. (2001) Mechanisms of tryptophan fluorescence shifts in

proteins. Biophysical Journal, 80(5), 2093-2109.

Walther, F. J., Gordon, L. M. & Waring, A. J. (2016) Design of surfactant protein B peptide

mimics based on the saposin fold for synthetic lung surfactants. Biomedicine Hub,

1(3), 1-21.

Wang, Y., Grabowski, G. A. & Qi, X. (2003) Phospholipid vesicle fusion induced by

saposin C. Archives of Biochemistry and Biophysics, 415(1), 43-53.

Willis, C., Wang, C. K., Osman, A., Simon, A., Pickering, D., Mulvenna, J., Riboldi-

Tunicliffe, A., Jones, M. K., Loukas, A. & Hofmann, A. (2011) Insights into the

152

membrane interactions of the saposin-like proteins Na-SLP-1 and Ac-SLP-1 from

human and dog hookworm. PLoS ONE, 6(10), e25369.

Yatsu, L. Y. & Jacks, T. J. (1968) Association of lysosomal activity with aleurone grains

in plant seeds. Archives of Biochemistry and Biophysics, 124, 466-471.

Yomo, H. & Srinivasan, K. (1973) Protein breakdown and formation of protease in

attached and detached cotyledons of Phaseolus vulgaris L. Plant Physiology, 52(6),

671-673.

153

4.6 Appendices

4.6-Appendix A Comparison of the secondary structures of StAP, phytepsin, cardosin

A and AtAP PSIs at different pH in iso-ionic buffered saline.

Figure 4.8: Far-UV CD spectra comparing PSIs at different pH in iso-ionic buffered saline.

Spectra are arranged for comparison between pH values for the respective PSIs

154

4.6-Appendix B Comparison of the secondary structures of StAP, phytepsin, cardosin

A and AtAP PSIs at different pH in iso-ionic buffered saline

Figure 4.9: Far-UV CD spectra comparing PSIs at different pH in iso-ionic buffered

saline. Spectra are arranged for comparison between PSIs at given pH values

155

Chapter 5: Concluding Discussion

The experiments of the previous three chapters comprised a series of related protein

biophysical and biochemical studies that sought to gain insight into the structural basis of

swaposin bilayer activity. From the beginning, the overriding focus was to identify

structural elements that govern PSI-bilayer interactions. The findings represent a

significant improvement to the current state of knowledge concerning the mode of action

of the poorly understood and structurally unique plant-derived saposin-like AP domains.

5.1 Further Considerations for PSI-Induced Bilayer Fusion

The use of DLS in conjunction with electron microscopy has long been recognized as

reliable for detecting vesicle fusion (Day et al, 1977). Additionally, the use of DLS for

detecting surface active protein association with vesicles has been used up to and including

a recent study on an amoeboid saposin-like protein (Michalek & Leippe, 2015) as well as

for SNARE-induced membrane fusion (Brüning et al, 2013; Castorph et al, 2011; Trivedi

et al, 2000; Versluis et al, 2014; Yang et al, 2015). Although the technique has long proven

to be useful for assessing vesicle size changes in solution in a practical manner, it is not

capable of giving morphological information for LUVs under the required assay

conditions. In fact, measuring liposome size by DLS for the assays explicitly assumes that

size changes result in approximately spherical particles. Although DLS is capable of

assessing changes in particle shape for other experimental setups, tracking liposome size

increases requires that the speed of measurements is a priority over precision.

The expectation to visualize PSI-induced changes on bilayers by cryo-TEM initially

resulted in empty fields of view or inconsistent and distorted images. This perhaps was

156

fortuitous because it induced added scepticism for images that would follow, and further

consideration of potential experimental pitfalls. After experimental adjustments, various

shapes were visible for PSI-treated vesicles (Chapter 3, Figure 3.8), many of which were

drastic deviations from control vesicles. It had been anticipated that ~100 nm vesicles

would be observed, mixed with an array of approximately spherical liposomes, some

several times their original size. This expectation was based on apparent size as measured

by DLS, the only indication, to that point, of fusion product changes by PSIs on bilayer

vesicles (see Chapter 2, Table 2.3). Initially, high quality images were selected for their

contrast and sharpness as well as exemplifying the drastic shape changes between control

and test conditions. In re-assessing all images that had been collected for the cryo-TEM

experiments, it was subsequently determined that some of the images having superior

contrast were in fact showing vesicles that were frozen on the grid support itself as opposed

to vesicles in solution frozen within the holes of the grid matrix as desired. This situation

causes image distortions due to mobility of material on grid surfaces (as evidenced by

streaking and/or odd elongated forms), resulting in unreliable sizes and shapes of mobile

vesicles. Additionally, interaction with the grid surface results in artefacts arising from

clumping/aggregation of vesicles. This phenomenon is not indicative of true vesicle

characteristics regardless of test or control conditions. Phenomena observed for vesicles

contacting grid surfaces cannot be considered as reliable. Illustrating this point, a high

quality cryo-TEM image selected by the manufacturer of cryo-TEM imaging hardware and

software shows clumping and distortion of vesicles on grid surfaces including contact with

the edges of the pore/intra-grid space, while those properly situated within the pore showed

uniformity of size and spherical shape (Goodwin & Khant).

157

Limiting consideration to vesicles not in contact with grid surfaces, the results of the

cryo-TEM experiments were outlined and discussed previously in Chapter 3, Figures 3.8

and 3.9. Although it remains premature to assert a PSI-induced bilayer disruption scheme

with confidence, the cryo-TEM results of Chapter 3 produced vesicle morphologies that

were repeated using different samples on different days and in different

concentrations/ratios of components. In the light of this reproducibility as well as the

context provided by published reports for other bilayer/surface-active proteins/peptides

(see the Discussion section of Chapter 3), it seems reasonable to postulate that the observed

vesicle morphologies may constitute a coherent sequence of occurrence. It should be noted

that the following arrangement is intended strictly as an exploratory supposition which can

serve as a reference point for future imaging interpretation and comparison. In Figure 5.1,

various distinct liposome shapes from the original cryo-TEM images (Chapter 3, Figure

3.8) have been arranged in a fashion that appears to reveal a progression of morphologies

from intact spherical vesicles (upper left) to narrow rod-like objects having a clear lipid

bilayer surrounding a dense core (lower left). When considering the scheme in Figure 5.1,

consider that quantitation of the various morphologies and their sizes were presented in

Chapter 3, including increased size range as well as overall size among the spherical vesicle

population, suggesting fusion of spherical LUVs. Thus, the purported progression in Figure

5.1 would either represent phenomena separate from spherical vesicle growth, or possibly

the latter constitutes a part of the progression in morphological change. At present, it

remains for future investigations to systematically determine morphological changes to

PSI-treated bilayers under an array of conditions and concentrations, and significant access

to cryo-TEM facilities and technical expertise.

158

Figure 5.1: Morphologies of PSI-treated LUVs observed by cryo-TEM. The selected

shapes are arranged to illustrate an apparent progression of morphological changes.

159

5.2 Bilayer Activity Via an Isolated Swaposin Structural Region

As discussed in Section 3.5 of Chapter 3, the singular portion of the PSI sub-structure

that displayed activity of its own was the penultimate helix, H3, located in the C-terminal

half of the overall primary structure. Recall the features of H3:

EAPLCTACEMAVVWMQNQLKQ

That a sub-structure such as H3 was found to be “active” was not surprising in and of itself;

in fact, it was hypothesized that the adjacent helix region would display membrane

perturbation activity. A similar case for saposin was previously reported (Wang et al,

2003), albeit not in the same mode as the present case with respect to sequence.

Furthermore, the fact that H3 is active, and Lys83 is essential for H3 bilayer activity, has

now offered a simple and fortuitous case study to consider in future experiments. This is

particularly true in consideration that the [Asn/Gln]-[Asn/Gln]-[Ala/Leu/Ile/Val]-

[Arg/Lys]-[Asn/Gln] motif, found in H3, is present in many flocculating/coagulating

proteins. Flocculating proteins such as Flo must have a positively charged, glutamine-rich

portion of the peptide in order to cause flocculation (Suarez et al, 2005). There are

variations on the general structural arrangement of having high Gln content, a hydrophobic

moiety and at least one positive charge (Lys or Arg) within a relatively short sequence. To

the best of our knowledge, the shortest such example that functions as a flocculent protein

is β-lactoglobulin fragment 1−8 (Pouliot et al, 2009). Other examples of similar

antimicrobial peptides/proteins are listed in Table 3.1.

As discussed in Chapter 3, StAP PSI contains an additional Gln (QNQLKQ)

compared to the 2S/saposin 5-residue motif above. The fact that there appears to be just 28

non-redundant protein structures known, listed in the table in Table 3.2 for reference, attests

160

to its rarity. Furthermore, the apparent commonality of function among all of the known

28 structures surely points to a critical role for this rare motif, and a functionality that

likewise presumably is manifested in H3. The seeming differences in modes of action

between full PSI and H3, discussed in Chapter 3, may indicate that the functionality of the

H3 sequence is available exclusively upon its structural availability for interaction, i.e.,

steric and/or other conditions that hinder or prevent bilayer- and/or protein-protein

interactions. If this is true then presumably the presence of the rare motif, as opposed to a

general abundance of amphiphilic character and Asn/Gln residues typical for antimicrobial

sequences (Patrzykat & Douglas, 2005), is a chance event since there are no reports of PSI

sequence fragments present in vivo. Also, unlike StAP PSI, the motif is not present in the

three other PSIs from Chapter 4 despite strict sequence conservation among the four

species at multiple sequence portions (see Figure 4.1).

Irrespective of in vivo roles, the presence of the unique motif in a subdivision of the

PSI structure that has now been shown to target bilayers is a situation that deserves further

investigation directly on H3 variants by residue and/or motif swapping, based on

bioinformatics analyses of an array of species. Gaining insight into the variability among

different PSI species will provide context to phenomena observed in the present work

including the significance and occurrence of the QNQLKQ sequence among PSI species.

As of this writing, knowledge and understanding of biochemical variability among

SAPLIPS continues to exist mainly from the perspective of non-plant examples with the

exception of StAP PSI, a situation that restrains predicting connections between measured

functional differences and observed structural features. The focus of studies in the

161

immediate future should therefore include a comparative approach for a range of plant

SAPLIPs to illuminate the structure-function landscape in which StAP PSI exists.

5.3 Considerations for in vitro Structure-Function Studies on Bioactive Proteins

By necessity, protein structure-function experiments rely on techniques that employ

indirect means of detecting and measuring phenomena that exist on a sub-microscopic

scale. Critical for techniques that provide indirect sensing of sample states/characteristics

is a consistent consideration of the assumptions that underpin interpretations of observed

phenomena. Minimally complex, highly predictable assay systems are the ideal

experimental environment in which to conduct structure-function work, particularly for

research targets that have not seen significant attention. For this reason, the decision was

made to rely largely on a simple bilayer system consisting of 100 nm LUVs composed of

few phospholipid types. Importantly, the system provided reproducibility, familiarity in

terms of existing published findings, and reliability and confidence in observed test signals

(i.e., the vesicles did not spontaneously leak or aggregate/fuse, and quality of batch-to-

batch vesicle preparations was verifiable).

On the subject of PSI concentrations used throughout the various experiments, a range

of PSI/peptide concentrations was used depending on detection limits of the technique

and/or strength of leakage activity, mindful of remaining in a biologically-relevant range.

All concentrations fell within 0.5-10 µM with the exception of vesicle leakage screening

at higher concentrations in order to confirm the absence of activity for peptide/PSI in given

conditions (e.g., PSIs at neutral pH, or inactive peptides). The only other exception was the

use of 16 µM PSI in conjunction with elevated an phospholipid concentration, 1000 µM,

162

for cryo-TEM. The above range of PSI/peptide concentrations is well-aligned with

published data for StAP PSI antimicrobial activity (Muñoz et al, 2010) (see Table 5.1) as

well as a varied selection of other plant antibacterial/antifungal proteins. IC50 values of

plant antimicrobial peptides (thionins, plant defensins, LTP, and snakins (Broekaert et al,

1997; Segura et al, 1999) against bacterial and fungal pathogens generally range from 0.10

to 5.0 µM. Examples of plant antifungal peptides’ IC50 ranges against various fungal plant

pathogens relevant to the respective biological situations are summarized in Table 5.2.

Thus, the concentration range used in the present study is in line with the biologically-

relevant concentrations for a wide variety of plant defense peptides/proteins.

5.4 Final Thoughts on Future Research Directions

Knowledge gained from the investigations detailed herein regarding PSI structural

features (Chapter 2), apparent modular design for fusogenic and disruption activities

(Chapter 3), and comparison of bilayer functionality among four plant species (Chapter 4)

will hopefully contribute to initiating further swaposin structure-function investigations.

The identification of H3 as a high interest sequence presents an opportunity to scan not

only potato PSI sequence, but putative PSI sequences not yet confirmed. Although the

focus of Chapter 3 leaned to the N-terminal portion of PSI due to the activities of H3, the

pH-sensitive H1 and H1H2 regions appear to be involved in StAP PSI pH sensitivity which

in turn may contribute to the overall pH–dependence of dimerization, critical to PSI

activity. This same N-terminal region was identified as being structurally similar to

hemagglutinin fusion peptide (Chapter 2), and is expected to be involved in membrane

insertion based on its clear structural design as a fusion peptide. To shed light on the

163

Table 5.1: Biologically active concentrations for StAP PSI. Antimicrobial activities

were reported by (Muñoz et al, 2010).

Antimicrobial Activity IC50 IC90 MBC

Inhibition of Fusarium solani spore germination 1.3 µM 10 µM -

Inhibition of Phytophthora infestans spores 0.25 µM 0.70 µM -

Inhibition of Staphylococcus aureus - - 21.0 µM

Inhibition of Bacillus cereus - - 3.5 µM

Inhibition of Escherichia coli - - 3.8 µM

IC: Inhibitory concentration, MBC: Minimal Bactericidal Concentration

Table 5.2: Concentration ranges of antimicrobial peptides used in vitro.

Antifungal peptide Concentration Range Reference

Histatin 1.4 µM Kavanagh and Dowd (2004)

Melittin 6.0 – 25.0 µM López-García et al. (2000)

Synthetic hexapeptide 66-10 10.0 – 45.0 µM López-García et al. (2000)

Cecropin-derived peptide 2 6.5 – 30.0 µM Cavallarin et al. (1998)

French Bean Defensin-Like

Antifungal Peptide 3.0 – 4.5 µM Leung et al. (2008)

164

energetics and dynamics of the PSI Trp77 transition from protein-protein to protein-lipid

contacts, in silico experiments should be designed and carried out. Understanding the

factors that govern sequestering of hydrophobic residues within monomeric PSI at neutral

(inactive) pH as well as within the dimer hydrophobic core in acidic conditions, and the

possible subsequent transition to the bilayer hydrophobic interior is likely key for

unravelling the detailed mechanism for swaposin bilayer interaction and perturbation.

In the near future, it seems reasonable to expect an expansion of scientific enquiry

regarding plant saposins from a biotechnological standpoint, partly because of their

possible uses in recombinant technologies as antifungal agents and/or intelligent crop breed

selection as a desirable trait for plant pathogen resistance. Two key studies regarding in

vitro antimicrobial activities of potato PSI, including human pathogens, made it clear that

plant saposins are potentially useful biotechnologically (Mendieta et al, 2006; Muñoz et al,

2010). At present, saposins are already fruitful in terms of wide-reaching and potentially

impactful applications (Fonseca et al, 2009; Frauenfeld et al, 2016; Kaimal et al, 2011; Qi,

2010; 2012; Qi et al, 2009; Walther et al, 2016). The plant kingdom is not only potentially

a near limitless source of biochemical variability, it can also be a source of evolutionarily-

ancient template sequences and structures. One would expect expansion into the plant

saposin realm by researchers with envisioned saposin-based biotechnological designs,

seeking a broader base of natural templates and exotic or possibly unique motifs and/or

activities. Exploiting the variability between highly similar plant saposin family members

offers the potential to gain a fundamental understanding of how plants employ these critical

defense proteins for different purposes/situations. In turn, such knowledge will enable

predictive structure redesign / engineering based upon the hundreds, if not thousands, of

165

putative PSI gene sequences already stored in the world’s plant genome projects, ready to

serve as templates for understanding the roles of plant SAPLIP in issues related to food

security, agricultural disease control and sustainability, and possibly human disease

treatment.

166

5.5 References

Broekaert, W. F., Cammue, B. P. A., De Bolle, M. F. C., Thevissen, K., De Samblanx, G.

W., Osborn, R. W. & Nielson, K. (1997) Antimicrobial peptides from plants. Critical

Reviews in Plant Sciences, 16(3), 297-323.

Brüning, B., Stehle, R., Falus, P. & Farago, B. (2013) Influence of charge density on bilayer

bending rigidity in lipid vesicles: A combined dynamic light scattering and neutron spin-

echo study. The European Physical Journal E, 36(7), 1-8.

Castorph, S., Henriques, S. S., Holt, M., Riedel, D., Jahn, R. & Salditt, T. (2011) Synaptic

vesicles studied by dynamic light scattering. The European Physical Journal E, 34(6),

1-11.

Cavallarin, L., Andreu, D. & San Segundo, B. (1998) Cecropin A—derived peptides are

potent inhibitors of fungal plant pathogens. Molecular Plant-Microbe Interactions,

11(3), 218-227.

Day, E. P., Ho, J. T., Kunze, R. K. & Sun, S. T. (1977) Dynamic light scattering study of

calcium-induced fusion in phospholipid vesicles. Biochimica et Biophysica Acta (BBA)-

Biomembranes, 470(3), 503-508.

Fonseca, S. B., Pereira, M. P. & Kelley, S. O. (2009) Recent advances in the use of cell-

penetrating peptides for medical and biological applications. Advanced Drug Delivery

Reviews, 61(11), 953-964.

Frauenfeld, J., Loving, R., Armache, J.-P., Sonnen, A. F. P., Guettou, F., Moberg, P., Zhu,

L., Jegerschold, C., Flayhan, A., Briggs, J. A. G., Garoff, H., Low, C., Cheng, Y. &

Nordlund, P. (2016) A saposin-lipoprotein nanoparticle system for membrane proteins.

Nature Methods, 13(4), 345-351.

Goodwin, J. & Khant, H. Frozen hydrated lipid vesicles. http://www.gatan.com/-

products/tem-specimen-preparation/cryoplunge-3-system#publication-related.

National Center for Macromolecular Imaging, Baylor College of Medicine, Houston,

TX, USA.

Kaimal, V., Chu, Z., Mahller, Y. Y., Papahadjopoulos-Sternberg, B., Cripe, T. P., Holland,

S. K. & Qi, X. (2011) Saposin C Coupled Lipid Nanovesicles Enable Cancer-Selective

167

Optical and Magnetic Resonance Imaging. Molecular Imaging and Biology, 13(5), 886-

897.

Kavanagh, K. & Dowd, S. (2004) Histatins: Antimicrobial peptides with therapeutic

potential. Journal of Pharmacy and Pharmacology, 56(3), 285-289.

Leung, E. H. W., Wong, J. H. & Ng, T. B. (2008) Concurrent purification of two defense

proteins from French bean seeds: A defensin-like antifungal peptide and a

hemagglutinin. Journal of Peptide Science, 14(3), 349-353.

López-García, B., González-Candelas, L., Pérez-Payá, E. & Marcos, J. F. (2000)

Identification and characterization of a hexapeptide with activity against

phytopathogenic fungi that cause postharvest decay in fruits. Molecular Plant-Microbe

Interactions, 13(8), 837-846.

Mendieta, J. R., Pagano, M. R., Munoz, F. F., Daleo, G. R. & Guevara, M. G. (2006)

Antimicrobial activity of potato aspartic proteases (StAPs) involves membrane

permeabilization. Microbiology, 152, 2039-2047.

Michalek, M. & Leippe, M. (2015) Mechanistic insights into the lipid interaction of an

ancient saposin-like protein. Biochemistry, 54(9), 1778-1786.

Muñoz, F. F., Mendieta, J. R., Pagano, M. R., Paggi, R. A., Daleo, G. R. & Guevara, M.

G. (2010) The swaposin-like domain of potato aspartic protease (StAsp-PSI) exerts

antimicrobial activity on plant and human pathogens. Peptides, 31(5), 777-785.

Patrzykat, A. & Douglas, S. E. (2005) Antimicrobial peptides: Cooperative approaches to

protection. Protein and Peptide Letters, 12(1), 19-25.

Pouliot, Y., Guy, M.-M., Tremblay, M., Gaonac’h, A.-C., Chay Pak Ting, B. P., Gauthier,

S. F. & Voyer, N. (2009) Isolation and characterization of an aggregating peptide from

a tryptic hydrolysate of whey proteins. Journal of Agricultural and Food Chemistry,

57(9), 3760-3764.

Qi, X. (2010) Saposin C-DOPS: a novel anti-tumor agent. Patent application number US

10/801,517.

Qi, X. (2012) Fusogenic properties of saposin c and related proteins and peptides for

application to transmembrane drug delivery systems. Patent application number US

13/127,630 .

168

Qi, X., Chu, Z., Mahller, Y. Y., Stringer, K. F., Witte, D. P. & Cripe, T. P. (2009) Cancer-

Selective Targeting and Cytotoxicity by Liposomal-Coupled Lysosomal Saposin C

Protein. Clinical Cancer Research, 15(18), 5840.

Segura, A., Moreno, M., Madueño, F., Molina, A. & García-Olmedo, F. (1999) Snakin-1,

a Peptide from Potato That Is Active Against Plant Pathogens. Molecular Plant-Microbe

Interactions, 12(1), 16-23.

Suarez, M., Haenni, M., Canarelli, S., Fisch, F., Chodanowski, P., Servis, C., Michielin,

O., Freitag, R., Moreillon, P. & Mermod, N. (2005) Structure-Function Characterization

and Optimization of a Plant-Derived Antibacterial Peptide. Antimicrobial Agents and

Chemotherapy, 49(9), 3847-3857.

Trivedi, V. D., Yu, C., Veeramuthu, B., Francis, S. & Chang, D. K. (2000) Fusion induced

aggregation of model vesicles studied by dynamic and static light scattering. Chemistry

and Physics of Lipids, 107(1), 99-106.

Versluis, F., Voskuhl, J., Vos, J., Friedrich, H., Ravoo, B. J., Bomans, P. H. H., Stuart, M.

C. A., Sommerdijk, N. A. J. M. & Kros, A. (2014) Coiled coil driven membrane fusion

between cyclodextrin vesicles and liposomes. Soft Matter, 10(48), 9746-9751.

Walther, F. J., Gordon, L. M. & Waring, A. J. (2016) Design of surfactant protein B peptide

mimics based on the saposin fold for synthetic lung surfactants. Biomedicine Hub, 1(3),

1-21.

Wang, Y., Grabowski, G. A. & Qi, X. (2003) Phospholipid vesicle fusion induced by

saposin C. Archives of Biochemistry and Biophysics, 415(1), 43-53.

Yang, Y., Heo, P., Kong, B., Park, J.-B., Jung, Y.-H., Shin, J., Jeong, C. & Kweon, D.-H.

(2015) Dynamic light scattering analysis of SNARE-driven membrane fusion and the

effects of SNARE-binding flavonoids. Biochemical and Biophysical Research

Communications, 465(4), 864-870.